Stephen Hawking
A Brief History of Time
Commemorative Coin

One side of this Silver Plated Coin has an image of a galaxy in space with the words "ABriefer History on Time" with 2 stars

The back has an image of the great man with his signature and the years of his life 1942 - 2018

The coin is 40mm in diameter, weighs about 1 oz and it comes in air-tight acrylic coin holder.

in Excellent Condition.


Bid with Confidence - Check My 100% Positive Feedback from over 2,500 Satisfied Customers
I have over 10 years of Ebay Selling Experience - So Why Not Treat Yourself?

I have got married recently and need to raise funds to meet the costs also we are planning to move into a house together

 I always combined postage on multiple items

Instant Feedback Automatically Left Immediately after Receiving Payment

All Items Sent out within 24 hours of Receiving Payment.

Overseas Bidders Please Note Surface Mail Delivery Times >

Western Europe takes up to 2 weeks, 

Eastern Europe up to 5 weeks, 

North America up to 6 weeks, 

South America, Africa and Asia up to 8 weeks and 

Australasia up to 12 weeks


Thanks for Looking and Best of Luck with the Bidding!!

Also if bidding from overseas and you want your item tracked please select the International Signed for Postage Option
For that Interesting Conversational Piece, A Birthday Present, Christmas Gift, A Comical Item to Cheer Someone Up or That Unique Perfect Gift for the Person Who has Everything....You Know Where to Look for a Bargain!

XXXX - DO NOT CLICK HERE - XXXX 

Click Here to Add me to Your List of Favourite Sellers

If You Have any Questions Please Message me through ebay and I Will Reply ASAP


have sold items to coutries such as Afghanistan * Albania * Algeria * American Samoa (US) * Andorra * Angola * Anguilla (GB) * Antigua and Barbuda * Argentina * Armenia * Aruba (NL) * Australia * Austria * Azerbaijan * Bahamas * Bahrain * Bangladesh * Barbados * Belarus * Belgium * Belize * Benin * Bermuda (GB) * Bhutan * Bolivia * Bonaire (NL)  * Bosnia and Herzegovina * Botswana * Bouvet Island (NO) * Brazil * British Indian Ocean Territory (GB) * British Virgin Islands (GB) * Brunei * Bulgaria * Burkina Faso * Burundi * Cambodia * Cameroon * Canada * Cape Verde * Cayman Islands (GB) * Central African Republic * Chad * Chile * China * Christmas Island (AU) * Cocos Islands (AU) * Colombia * Comoros * Congo * Democratic Republic of the Congo * Cook Islands (NZ) * Coral Sea Islands Territory (AU) * Costa Rica * Croatia * Cuba * Curaçao (NL)  * Cyprus * Czech Republic * Denmark * Djibouti * Dominica * Dominican Republic * East Timor * Ecuador * Egypt * El Salvador * Equatorial Guinea * Eritrea * Estonia * Ethiopia * Falkland Islands (GB) * Faroe Islands (DK) * Fiji Islands * Finland * France * French Guiana (FR) * French Polynesia (FR) * French Southern Lands (FR) * Gabon * Gambia * Georgia * Germany * Ghana * Gibraltar (GB) * Greece * Greenland (DK) * Grenada * Guadeloupe (FR) * Guam (US) * Guatemala * Guernsey (GB) * Guinea * Guinea-Bissau * Guyana * Haiti * Heard and McDonald Islands (AU) * Honduras * Hong Kong (CN) * Hungary * Iceland * India * Indonesia * Iran * Iraq * Ireland * Isle of Man (GB) * Israel * Italy * Ivory Coast * Jamaica * Jan Mayen (NO) * Japan * Jersey (GB) * Jordan * Kazakhstan * Kenya * Kiribati * Kosovo * Kuwait * Kyrgyzstan * Laos * Latvia * Lebanon * Lesotho * Liberia * Libya * Liechtenstein * Lithuania * Luxembourg * Macau (CN) * Macedonia * Madagascar * Malawi * Malaysia * Maldives * Mali * Malta * Marshall Islands * Martinique (FR) * Mauritania * Mauritius * Mayotte (FR) * Mexico * Micronesia * Moldova * Monaco * Mongolia * Montenegro * Montserrat (GB) * Morocco * Mozambique * Myanmar * Namibia * Nauru * Navassa (US) * Nepal * Netherlands * New Caledonia (FR) * New Zealand * Nicaragua * Niger * Nigeria * Niue (NZ) * Norfolk Island (AU) * North Korea * Northern Cyprus * Northern Mariana Islands (US) * Norway * Oman * Pakistan * Palau * Palestinian Authority * Panama * Papua New Guinea * Paraguay * Peru * Philippines * Pitcairn Island (GB) * Poland * Portugal * Puerto Rico (US) * Qatar * Reunion (FR) * Romania * Russia * Rwanda * Saba (NL)  * Saint Barthelemy (FR) * Saint Helena (GB) * Saint Kitts and Nevis * Saint Lucia * Saint Martin (FR) * Saint Pierre and Miquelon (FR) * Saint Vincent and the Grenadines * Samoa * San Marino * Sao Tome and Principe * Saudi Arabia * Senegal * Serbia * Seychelles * Sierra Leone * Singapore * Sint Eustatius (NL)  * Sint Maarten (NL)  * Slovakia * Slovenia * Solomon Islands * Somalia * South Africa * South Georgia (GB) * South Korea * South Sudan * Spain * Sri Lanka * Sudan * Suriname * Svalbard (NO) * Swaziland * Sweden * Switzerland * Syria * Taiwan * Tajikistan * Tanzania * Thailand * Togo * Tokelau (NZ) * Tonga * Trinidad and Tobago * Tunisia * Turkey * Turkmenistan * Turks and Caicos Islands (GB) * Tuvalu * U.S. Minor Pacific Islands (US) * U.S. Virgin Islands (US) * Uganda * Ukraine * United Arab Emirates * United Kingdom * United States * Uruguay * Uzbekistan * Vanuatu * Vatican City * Venezuela * Vietnam * Wallis and Futuna (FR) * Yemen * Zambia * Zimbabwe and major cities such as Tokyo, Yokohama, New York City, Sao Paulo, Seoul, Mexico City, Osaka, Kobe, Kyoto, Manila, Mumbai, Delhi, Jakarta, Lagos, Kolkata, Cairo, Los Angeles, Buenos Aires, Rio de Janeiro, Moscow, Shanghai, Karachi, Paris, Istanbul, Nagoya, Beijing, Chicago, London, Shenzhen, Essen, Düsseldorf, Tehran, Bogota, Lima, Bangkok, Johannesburg, East Rand, Chennai, Taipei, Baghdad, Santiago, Bangalore, Hyderabad, St Petersburg, Philadelphia, Lahore, Kinshasa, Miami, Ho Chi Minh City, Madrid, Tianjin, Kuala Lumpur, Toronto, Milan, Shenyang, Dallas, Fort Worth, Boston, Belo Horizonte, Khartoum, Riyadh, Singapore, Washington, Detroit, Barcelona,, Houston, Athens, Berlin, Sydney, Atlanta, Guadalajara, San Francisco, Oakland, Montreal, Monterey, Melbourne, Ankara, Recife, Phoenix/Mesa, Durban, Porto Alegre, Dalian, Jeddah, Seattle, Cape Town, San Diego, Fortaleza, Curitiba, Rome, Naples, Minneapolis, St. Paul, Tel Aviv, Birmingham, Frankfurt, Lisbon, Manchester, San Juan, Katowice, Tashkent, Fukuoka, Baku, Sumqayit, St. Louis, Baltimore, Sapporo, Tampa, St. Petersburg, Taichung, Warsaw, Denver, Cologne, Bonn, Hamburg, Dubai, Pretoria, Vancouver, Beirut, Budapest, Cleveland, Pittsburgh, Campinas, Harare, Brasilia, Kuwait, Munich, Portland, Brussels, Vienna, San Jose, Damman , Copenhagen, Brisbane, Riverside, San Bernardino, Cincinnati and Accra 3


Professor Stephen Hawking 1942-2018 Credit: Andre Pattenden Friends and colleagues from the University of Cambridge have paid tribute to Professor Stephen Hawking, who died today at the age of 76. Widely regarded as one of the world’s most brilliant minds, he was known throughout the world for his contributions to science, his books, his television appearances, his lectures and through biographical films. He leaves three children and three grandchildren. Professor Hawking broke new ground on the basic laws which govern the universe, including the revelation that black holes have a temperature and produce radiation, now known as Hawking radiation. At the same time, he also sought to explain many of these complex scientific ideas to a wider audience through popular books, most notably his bestseller A Brief History of Time. He was awarded the CBE in 1982, was made a Companion of Honour in 1989, and was awarded the US Presidential Medal of Freedom in 2009. He was the recipient of numerous awards, medals and prizes, including the Copley Medal of the Royal Society, the Albert Einstein Award, the Gold Medal of the Royal Astronomical Society, the Fundamental Physics Prize, and the BBVA Foundation Frontiers of Knowledge Award for Basic Sciences. He was a Fellow of The Royal Society, a Member of the Pontifical Academy of Sciences, and a Member of the US National Academy of Sciences. He achieved all this despite a decades-long battle with motor neurone disease, with which he was diagnosed while a student, and eventually led to him being confined to a wheelchair and to communicating via his instantly recognisable computerised voice. His determination in battling with his condition made him a champion for those with a disability around the world. Professor Hawking came to Cambridge in 1962 as a PhD student, and rose to become the Lucasian Professor of Mathematics, a position once held by Isaac Newton, in 1979. In 2009, he retired from this position and was the Dennis Stanton Avery and Sally Tsui Wong-Avery Director of Research in the Department of Applied Mathematics and Theoretical Physics until his death. He was also a member of the University's Centre for Theoretical Cosmology, which he founded in 2007. He was active scientifically and in the media until the end of his life. Professor Stephen Toope, Vice-Chancellor of the University of Cambridge, paid tribute, saying, “Professor Hawking was a unique individual who will be remembered with warmth and affection not only in Cambridge but all over the world. His exceptional contributions to scientific knowledge and the popularisation of science and mathematics have left an indelible legacy. His character was an inspiration to millions. He will be much missed.” Stephen William Hawking was born on January 8, 1942 in Oxford although his family was living in north London at the time. In 1959, the family moved to St Albans where he attended St Albans School. Despite the fact that he was always ranked at the lower end of his class by teachers, his school friends nicknamed him ‘Einstein’ and seemed to have encouraged his interest in science. In his own words, “physics and astronomy offered the hope of understanding where we came from and why we are here. I wanted to fathom the depths of the Universe.” His ambition brought him a scholarship to University College Oxford to read Natural Science. There he studied physics and graduated with a first class honours degree. He then moved to Trinity Hall, Cambridge and was supervised by Dennis Sciama at the Department of Applied Mathematics and Theoretical Physics for his PhD; his thesis was titled Properties of Expanding Universes. In 2017, he made his PhD thesis freely available online via the University of Cambridge’s Open Access repository. There have been over a million attempts to download the thesis, demonstrating the enduring popularity of Hawking and his academic legacy. On completion of his PhD Hawking became a research fellow at Gonville and Caius College where he remained a fellow for the rest of his life. During his early years at Cambridge, he was influenced by Roger Penrose and developed the singularity theorems which show that the Universe began with the Big Bang. An interest in singularities naturally led to an interest in black holes and his subsequent work in this area laid the foundations for the modern understanding of black holes. He proved that when black holes merge, the surface area of the final black hole must exceed the sum of the areas of the initial black holes, and he showed that this places limits on the amount of energy that can be carried away by gravitational waves in such a merger. He found that there were parallels to be drawn between the laws of thermodynamics and the behaviour of black holes. This eventually led, in 1974, to the revelation that black holes have a temperature and produce radiation, now known as Hawking radiation, a discovery which revolutionised theoretical physics. He also realised that black holes must have an entropy – often described as a measure of how much disorder is present in a given system – equal to one quarter of the area of their event horizon: – the ‘point of no return’, where the gravitational pull of a black hole becomes so strong that escape is impossible. Some forty odd years later, the precise nature of this entropy is still a puzzle. However, these discoveries led to Hawking formulating the ‘information paradox’ which illustrates a fundamental conflict between quantum mechanics and our understanding of gravitational physics. This is probably the greatest mystery facing theoretical physicists today. To understand black holes and cosmology requires one to develop a theory of quantum gravity. Quantum gravity is an unfinished project which is attempting to unify general relativity, the theory of gravitation and of space and time with the ideas of quantum mechanics. Hawking’s work on black holes started a new chapter in this quest and most of his subsequent achievements centred on these ideas. Hawking recognised that quantum mechanical effects in the very early universe might provide the primordial gravitational seeds around which galaxies and other large-scale structures could later form. This theory of inflationary fluctuations, developed along with others in the early 1980s, is now supported by strong experimental evidence from the COBE, WMAP and Planck satellite observations of the cosmic microwave sky. Another influential idea was Hawking’s ‘no boundary’ proposal which resulted from the application of quantum mechanics to the entire universe. This idea allows one to explain the creation of the universe in a way that is compatible with laws of physics as we currently understand them. Professor Hawking’s influential books included The Large Scale Structure of Spacetime, with G F R Ellis; General Relativity: an Einstein centenary survey, with W Israel; Superspace and Supergravity, with M Rocek (1981); The Very Early Universe, with G Gibbons and S Siklos, and 300 Years of Gravitation, with W Israel. However, it was his popular science books which took Professor Hawking beyond the academic world and made him a household name. The first of these, A Brief History of Time, was published in 1988 and became a surprise bestseller, remaining on the Sunday Times best-seller list for a record-breaking 237 weeks. Later popular books included Black Holes and Baby Universes, The Universe in a Nutshell, A Briefer History of Time, and My Brief History. He also collaborated with his daughter Lucy on a series of books for children about a character named George who has adventures in space. In 2014, a film of his life, The Theory of Everything, was released. Based on the book by his first wife Jane, the film follows the story of their life together, from first meeting in Cambridge in 1964, with his subsequent academic successes and his increasing disability. The film was met with worldwide acclaim and Eddie Redmayne, who played Stephen Hawking, won the Academy Award for Best Actor at the 2015 ceremony. Travel was one of Professor Hawking’s pastimes. One of his first adventures was to be caught up in the 7.1 magnitude Bou-in-Zahra earthquake in Iran in 1962. In 1997 he visited the Antarctic. He has plumbed the depths in a submarine and in 2007 he experienced weightlessness during a zero-gravity flight, routine training for astronauts. On his return to ground he quipped “Space, here I come.” Writing years later on his website, Professor Hawking said: “I have had motor neurone disease for practically all my adult life. Yet it has not prevented me from having a very attractive family and being successful in my work. I have been lucky that my condition has progressed more slowly than is often the case. But it shows that one need not lose hope.” At a conference In Cambridge held in celebration of his 75th birthday in 2017, Professor Hawking said “It has been a glorious time to be alive and doing research into theoretical physics. Our picture of the Universe has changed a great deal in the last 50 years, and I’m happy if I’ve made a small contribution.” And he said he wanted others to feel the passion he has for understanding the universal laws that govern us all. “I want to share my excitement and enthusiasm about this quest. So remember to look up at the stars and not down at your feet. Try to make sense of what you see and wonder about what makes the universe exist. Be curious, and however difficult life may seem, there is always something you can do, and succeed at. It matters that you don’t just give up.” Words: Tom Kirk, Sarah Collins Images: Alan Fersht, Graham CopeKoga, Andre Pattenden, Sir Cam, Dan White Universe Article Talk Read View source View history Tools This is a good article. Click here for more information. Page semi-protected Listen to this article From Wikipedia, the free encyclopedia For other uses, see Universe (disambiguation). Universe The Hubble Ultra-Deep Field image shows some of the most remote galaxies visible to present technology (diagonal is ~1/10 apparent Moon diameter)[1] Age (within ΛCDM model) 13.787 ± 0.020 billion years[2] Diameter Unknown.[3] Observable universe: 8.8×1026 m (28.5 Gpc or 93 Gly)[4] Mass (ordinary matter) At least 1053 kg[5] Average density (with energy) 9.9×10−27 kg/m3[6] Average temperature 2.72548 K (−270.4 °C, −454.8 °F)[7] Main contents Ordinary (baryonic) matter (4.9%) Dark matter (26.8%) Dark energy (68.3%)[8] Shape Flat with 4‰ error margin[9] The universe is all of space and time[a] and their contents.[10] It comprises all of existence, any fundamental interaction, physical process and physical constant, and therefore all forms of energy and matter, and the structures they form, from sub-atomic particles to entire galaxies. Space and time, according to the prevailing cosmological theory of the Big Bang, emerged together 13.787±0.020 billion years ago,[11] and the universe has been expanding ever since. Today the universe has expanded into an age and size that is physically only in parts observable as the observable universe, which is approximately 93 billion light-years in diameter at the present day, while the spatial size, if any, of the entire universe is unknown.[3] Some of the earliest cosmological models of the universe were developed by ancient Greek and Indian philosophers and were geocentric, placing Earth at the center.[12][13] Over the centuries, more precise astronomical observations led Nicolaus Copernicus to develop the heliocentric model with the Sun at the center of the Solar System. In developing the law of universal gravitation, Isaac Newton built upon Copernicus's work as well as Johannes Kepler's laws of planetary motion and observations by Tycho Brahe. Further observational improvements led to the realization that the Sun is one of a few hundred billion stars in the Milky Way, which is one of a few hundred billion galaxies in the observable universe. Many of the stars in a galaxy have planets. At the largest scale, galaxies are distributed uniformly and the same in all directions, meaning that the universe has neither an edge nor a center. At smaller scales, galaxies are distributed in clusters and superclusters which form immense filaments and voids in space, creating a vast foam-like structure.[14] Discoveries in the early 20th century have suggested that the universe had a beginning and has been expanding since then.[15] According to the Big Bang theory, the energy and matter initially present have become less dense as the universe expanded. After an initial accelerated expansion called the inflationary epoch at around 10−32 seconds, and the separation of the four known fundamental forces, the universe gradually cooled and continued to expand, allowing the first subatomic particles and simple atoms to form. Giant clouds of hydrogen and helium were gradually drawn to the places where matter was most dense, forming the first galaxies, stars, and everything else seen today. From studying the effects of gravity on both matter and light, it has been discovered that the universe contains much more matter than is accounted for by visible objects; stars, galaxies, nebulas and interstellar gas. This unseen matter is known as dark matter,[16] (dark means that there is a wide range of strong indirect evidence that it exists, but we have not yet detected it directly) having come into existence alongside the rest of the physical universe before gradually gathering into a foam-like structure of filaments and voids and allowing other forms of matter to form together into visible structures. The ΛCDM model is the most widely accepted model of the universe. It suggests that about 69.2%±1.2% of the mass and energy in the universe is dark energy which is responsible for the acceleration of the expansion of the universe, and about 25.8%±1.1% is dark matter.[17] Ordinary ('baryonic') matter is therefore only 4.84%±0.1% of the physical universe.[17] Stars, planets, and visible gas clouds only form about 6% of the ordinary matter.[18] There are many competing hypotheses about the ultimate fate of the universe and about what, if anything, preceded the Big Bang, while other physicists and philosophers refuse to speculate, doubting that information about prior states will ever be accessible. Some physicists have suggested various multiverse hypotheses, in which the universe might be one among many.[3][19][20] Part of a series on Physical cosmology Big Bang · Universe Age of the universe Chronology of the universe Early universe Inflation · Nucleosynthesis Backgrounds Gravitational wave (GWB) Microwave (CMB) · Neutrino (CNB) Expansion · Future Hubble's law · Redshift Expansion of the universe FLRW metric · Friedmann equations Inhomogeneous cosmology Future of an expanding universe Ultimate fate of the universe Components · Structure Components Lambda-CDM model Dark energy · Dark matter Structure Shape of the universe Galaxy filament · Galaxy formation Large quasar group Large-scale structure Reionization · Structure formation Experiments Black Hole Initiative (BHI) BOOMERanG Cosmic Background Explorer (COBE) Dark Energy Survey Planck space observatory Sloan Digital Sky Survey (SDSS) 2dF Galaxy Redshift Survey ("2dF") Wilkinson Microwave Anisotropy Probe (WMAP) Scientists Aaronson Alfvén Alpher Copernicus de Sitter Dicke Ehlers Einstein Ellis Friedmann Galileo Gamow Guth Hawking Hubble Huygens Kepler Lemaître Mather Newton Penrose Penzias Rubin Schmidt Smoot Suntzeff Sunyaev Tolman Wilson Zeldovich List of cosmologists Subject history Discovery of cosmic microwave background radiation History of the Big Bang theory Timeline of cosmological theories Category Astronomy portal vte Definition Duration: 50 seconds.0:50 Hubble Space Telescope – Ultra-Deep Field galaxies to Legacy field zoom out (video 00:50; May 2, 2019) The physical universe is defined as all of space and time[a] (collectively referred to as spacetime) and their contents.[10] Such contents comprise all of energy in its various forms, including electromagnetic radiation and matter, and therefore planets, moons, stars, galaxies, and the contents of intergalactic space.[21][22][23] The universe also includes the physical laws that influence energy and matter, such as conservation laws, classical mechanics, and relativity.[24] The universe is often defined as "the totality of existence", or everything that exists, everything that has existed, and everything that will exist.[24] In fact, some philosophers and scientists support the inclusion of ideas and abstract concepts—such as mathematics and logic—in the definition of the universe.[26][27][28] The word universe may also refer to concepts such as the cosmos, the world, and nature.[29][30] Etymology The word universe derives from the Old French word univers, which in turn derives from the Latin word universus, meaning 'combined into one'.[31] The Latin word 'universum' was used by Cicero and later Latin authors in many of the same senses as the modern English word is used.[32] Synonyms A term for universe among the ancient Greek philosophers from Pythagoras onwards was τὸ πᾶν (tò pân) 'the all', defined as all matter and all space, and τὸ ὅλον (tò hólon) 'all things', which did not necessarily include the void.[33][34] Another synonym was ὁ κόσμος (ho kósmos) meaning 'the world, the cosmos'.[35] Synonyms are also found in Latin authors (totum, mundus, natura)[36] and survive in modern languages, e.g., the German words Das All, Weltall, and Natur for universe. The same synonyms are found in English, such as everything (as in the theory of everything), the cosmos (as in cosmology), the world (as in the many-worlds interpretation), and nature (as in natural laws or natural philosophy).[37] Chronology and the Big Bang Main articles: Big Bang and Chronology of the universe Nature timeline This box: viewtalkedit −13 — – −12 — – −11 — – −10 — – −9 — – −8 — – −7 — – −6 — – −5 — – −4 — – −3 — – −2 — – −1 — – 0 — Dark Ages Reionization Matter-dominated era Accelerated expansion Water on Earth Single-celled life Photosynthesis Multicellular life Vertebrates ← Earliest Universe ← Earliest stars ← Earliest galaxy ← Earliest quasar / black hole ← Omega Centauri ← Andromeda Galaxy ← Milky Way spirals ← NGC 188 star cluster ← Alpha Centauri ← Earth / Solar System ← Earliest known life ← Earliest oxygen ← Atmospheric oxygen ← Sexual reproduction ← Earliest fungi ← Earliest animals / plants ← Cambrian explosion ← Earliest mammals ← Earliest apes / humans L i f e (billion years ago) The prevailing model for the evolution of the universe is the Big Bang theory.[38][39] The Big Bang model states that the earliest state of the universe was an extremely hot and dense one, and that the universe subsequently expanded and cooled. The model is based on general relativity and on simplifying assumptions such as the homogeneity and isotropy of space. A version of the model with a cosmological constant (Lambda) and cold dark matter, known as the Lambda-CDM model, is the simplest model that provides a reasonably good account of various observations about the universe. The Big Bang model accounts for observations such as the correlation of distance and redshift of galaxies, the ratio of the number of hydrogen to helium atoms, and the microwave radiation background. In this schematic diagram, time passes from left to right, with the universe represented by a disk-shaped "slice" at any given time. Time and size are not to scale. To make the early stages visible, the time to the afterglow stage (really the first 0.003%) is stretched and the subsequent expansion (really by 1,100 times to the present) is largely suppressed. The initial hot, dense state is called the Planck epoch, a brief period extending from time zero to one Planck time unit of approximately 10−43 seconds. During the Planck epoch, all types of matter and all types of energy were concentrated into a dense state, and gravity—currently the weakest by far of the four known forces—is believed to have been as strong as the other fundamental forces, and all the forces may have been unified. The physics controlling this very early period (including quantum gravity in the Planck epoch) is not understood, so we cannot say what, if anything, happened before time zero. Since the Planck epoch, the universe has been expanding to its present scale, with a very short but intense period of cosmic inflation speculated to have occurred within the first 10−32 seconds.[40] This initial period of inflation would explain why space appears to be very flat, and is uniform on scales much larger than light could otherwise travel since the start of the universe. Within the first fraction of a second of the universe's existence, the four fundamental forces had separated. As the universe continued to cool from its inconceivably hot state, various types of subatomic particles were able to form in short periods of time known as the quark epoch, the hadron epoch, and the lepton epoch. Together, these epochs encompassed less than 10 seconds of time following the Big Bang. These elementary particles associated stably into ever larger combinations, including stable protons and neutrons, which then formed more complex atomic nuclei through nuclear fusion.[41][42] This process, known as Big Bang nucleosynthesis, lasted for about 17 minutes and ended about 20 minutes after the Big Bang, so only the fastest and simplest reactions occurred. About 25% of the protons and all the neutrons in the universe, by mass, were converted to helium, with small amounts of deuterium (a form of hydrogen) and traces of lithium. Any other element was only formed in very tiny quantities. The other 75% of the protons remained unaffected, as hydrogen nuclei.[41][42]: 27–42  After nucleosynthesis ended, the universe entered a period known as the photon epoch. During this period, the universe was still far too hot for matter to form neutral atoms, so it contained a hot, dense, foggy plasma of negatively charged electrons, neutral neutrinos and positive nuclei. After about 377,000 years, the universe had cooled enough that electrons and nuclei could form the first stable atoms. This is known as recombination for historical reasons; electrons and nuclei were combining for the first time. Unlike plasma, neutral atoms are transparent to many wavelengths of light, so for the first time the universe also became transparent. The photons released ("decoupled") when these atoms formed can still be seen today; they form the cosmic microwave background (CMB).[42]: 15–27  As the universe expands, the energy density of electromagnetic radiation decreases more quickly than does that of matter because the energy of each photon decreases as it is cosmologically redshifted. At around 47,000 years, the energy density of matter became larger than that of photons and neutrinos, and began to dominate the large scale behavior of the universe. This marked the end of the radiation-dominated era and the start of the matter-dominated era.[43]: 390  In the earliest stages of the universe, tiny fluctuations within the universe's density led to concentrations of dark matter gradually forming. Ordinary matter, attracted to these by gravity, formed large gas clouds and eventually, stars and galaxies, where the dark matter was most dense, and voids where it was least dense. After around 100–300 million years,[43]: 333  the first stars formed, known as Population III stars. These were probably very massive, luminous, non metallic and short-lived. They were responsible for the gradual reionization of the universe between about 200–500 million years and 1 billion years, and also for seeding the universe with elements heavier than helium, through stellar nucleosynthesis.[44] The universe also contains a mysterious energy—possibly a scalar field—called dark energy, the density of which does not change over time. After about 9.8 billion years, the universe had expanded sufficiently so that the density of matter was less than the density of dark energy, marking the beginning of the present dark-energy-dominated era.[45] In this era, the expansion of the universe is accelerating due to dark energy. Physical properties Main articles: Observable universe, Age of the universe, and Metric expansion of space Of the four fundamental interactions, gravitation is the dominant at astronomical length scales. Gravity's effects are cumulative; by contrast, the effects of positive and negative charges tend to cancel one another, making electromagnetism relatively insignificant on astronomical length scales. The remaining two interactions, the weak and strong nuclear forces, decline very rapidly with distance; their effects are confined mainly to sub-atomic length scales.[46]: 1470  The universe appears to have much more matter than antimatter, an asymmetry possibly related to the CP violation.[47] This imbalance between matter and antimatter is partially responsible for the existence of all matter existing today, since matter and antimatter, if equally produced at the Big Bang, would have completely annihilated each other and left only photons as a result of their interaction.[48] The universe also appears to have neither net momentum nor angular momentum, which absences follow[clarification needed] from accepted physical laws if the universe is finite. These laws are Gauss's law and the non-divergence of the stress–energy–momentum pseudotensor.[49] Size and regions See also: Observational cosmology Television signals broadcast from Earth will never reach the edges of this image. According to the general theory of relativity, far regions of space may never interact with ours even in the lifetime of the universe due to the finite speed of light and the ongoing expansion of space. For example, radio messages sent from Earth may never reach some regions of space, even if the universe were to exist forever: space may expand faster than light can traverse it.[50] The spatial region that can be observed with telescopes is called the observable universe, which depends on the location of the observer. The proper distance—the distance as would be measured at a specific time, including the present—between Earth and the edge of the observable universe is 46 billion light-years[51] (14 billion parsecs), making the diameter of the observable universe about 93 billion light-years (28 billion parsecs).[51] The distance the light from the edge of the observable universe has traveled is very close to the age of the universe times the speed of light, 13.8 billion light-years (4.2×109 pc), but this does not represent the distance at any given time because the edge of the observable universe and the Earth have since moved further apart.[52] For comparison, the diameter of a typical galaxy is 30,000 light-years (9,198 parsecs), and the typical distance between two neighboring galaxies is 3 million light-years (919.8 kiloparsecs).[53] As an example, the Milky Way is roughly 100,000–180,000 light-years in diameter,[54][55] and the nearest sister galaxy to the Milky Way, the Andromeda Galaxy, is located roughly 2.5 million light-years away.[56] Because humans cannot observe space beyond the edge of the observable universe, it is unknown whether the size of the universe in its totality is finite or infinite.[3][57][58] Estimates suggest that the whole universe, if finite, must be more than 250 times larger than a Hubble sphere.[59] Some disputed[60] estimates for the total size of the universe, if finite, reach as high as 10 10 10 122 {\displaystyle 10^{10^{10^{122}}}} megaparsecs, as implied by a suggested resolution of the No-Boundary Proposal.[61][b] Age and expansion Main articles: Age of the universe and Expansion of the universe Assuming that the Lambda-CDM model is correct, the measurements of the parameters using a variety of techniques by numerous experiments yield a best value of the age of the universe at 13.799 ± 0.021 billion years, as of 2015.[2] Astronomers have discovered stars in the Milky Way galaxy that are almost 13.6 billion years old. Over time, the universe and its contents have evolved. For example, the relative population of quasars and galaxies has changed[62] and the universe has expanded. This expansion is inferred from the observation that the light from distant galaxies has been redshifted, which implies that the galaxies are receding from us. Analyses of Type Ia supernovae indicate that the expansion is accelerating.[63][64] The more matter there is in the universe, the stronger the mutual gravitational pull of the matter. If the universe were too dense then it would re-collapse into a gravitational singularity. However, if the universe contained too little matter then the self-gravity would be too weak for astronomical structures, like galaxies or planets, to form. Since the Big Bang, the universe has expanded monotonically. Perhaps unsurprisingly, our universe has just the right mass–energy density, equivalent to about 5 protons per cubic metre, which has allowed it to expand for the last 13.8 billion years, giving time to form the universe as observed today.[65][66] There are dynamical forces acting on the particles in the universe which affect the expansion rate. Before 1998, it was expected that the expansion rate would be decreasing as time went on due to the influence of gravitational interactions in the universe; and thus there is an additional observable quantity in the universe called the deceleration parameter, which most cosmologists expected to be positive and related to the matter density of the universe. In 1998, the deceleration parameter was measured by two different groups to be negative, approximately −0.55, which technically implies that the second derivative of the cosmic scale factor a ¨ {\displaystyle {\ddot {a}}} has been positive in the last 5–6 billion years.[67][68] Spacetime Main articles: Spacetime and World line See also: Lorentz transformation Modern physics regards events as being organized into spacetime.[69] This idea originated with the special theory of relativity, which predicts that if one observer sees two events happening in different places at the same time, a second observer who is moving relative to the first will see those events happening at different times.[70]: 45–52  The two observers will disagree on the time T {\displaystyle T} between the events, and they will disagree about the distance D {\displaystyle D} separating the events, but they will agree on the speed of light c {\displaystyle c}, and they will measure the same value for the combination c 2 T 2 − D 2 {\displaystyle c^{2}T^{2}-D^{2}}.[70]: 80  The square root of the absolute value of this quantity is called the interval between the two events. The interval expresses how widely separated events are, not just in space or in time, but in the combined setting of spacetime.[70]: 84, 136 [71] The special theory of relativity cannot account for gravity. Its successor, the general theory of relativity, explains gravity by recognizing that spacetime is not fixed but instead dynamical. In general relativity, gravitational force is reimagined as curvature of spacetime. A curved path like an orbit is not the result of a force deflecting a body from an ideal straight-line path, but rather the body's attempt to fall freely through a background that is itself curved by the presence of other masses. A remark by John Archibald Wheeler that has become proverbial among physicists summarizes the theory: "Spacetime tells matter how to move; matter tells spacetime how to curve",[72][73] and therefore there is no point in considering one without the other.[15] The Newtonian theory of gravity is a good approximation to the predictions of general relativity when gravitational effects are weak and objects are moving slowly compared to the speed of light.[74]: 327 [75] The relation between matter distribution and spacetime curvature is given by the Einstein field equations, which require tensor calculus to express.[76]: 43 [77] The solutions to these equations include not only the spacetime of special relativity, Minkowski spacetime, but also Schwarzschild spacetimes, which describe black holes; FLRW spacetime, which describes an expanding universe; and more. The universe appears to be a smooth spacetime continuum consisting of three spatial dimensions and one temporal (time) dimension. Therefore, an event in the spacetime of the physical universe can therefore be identified by a set of four coordinates: (x, y, z, t). On average, space is observed to be very nearly flat (with a curvature close to zero), meaning that Euclidean geometry is empirically true with high accuracy throughout most of the universe.[78] Spacetime also appears to have a simply connected topology, in analogy with a sphere, at least on the length scale of the observable universe. However, present observations cannot exclude the possibilities that the universe has more dimensions (which is postulated by theories such as the string theory) and that its spacetime may have a multiply connected global topology, in analogy with the cylindrical or toroidal topologies of two-dimensional spaces.[79][80] Shape Main article: Shape of the universe The three possible options for the shape of the universe General relativity describes how spacetime is curved and bent by mass and energy (gravity). The topology or geometry of the universe includes both local geometry in the observable universe and global geometry. Cosmologists often work with a given space-like slice of spacetime called the comoving coordinates. The section of spacetime which can be observed is the backward light cone, which delimits the cosmological horizon. The cosmological horizon, also called the particle horizon or the light horizon, is the maximum distance from which particles can have traveled to the observer in the age of the universe. This horizon represents the boundary between the observable and the unobservable regions of the universe.[81][82] The existence, properties, and significance of a cosmological horizon depend on the particular cosmological model. An important parameter determining the future evolution of the universe theory is the density parameter, Omega (Ω), defined as the average matter density of the universe divided by a critical value of that density. This selects one of three possible geometries depending on whether Ω is equal to, less than, or greater than 1. These are called, respectively, the flat, open and closed universes.[83] Observations, including the Cosmic Background Explorer (COBE), Wilkinson Microwave Anisotropy Probe (WMAP), and Planck maps of the CMB, suggest that the universe is infinite in extent with a finite age, as described by the Friedmann–Lemaître–Robertson–Walker (FLRW) models.[84][79][85][86] These FLRW models thus support inflationary models and the standard model of cosmology, describing a flat, homogeneous universe presently dominated by dark matter and dark energy.[87][88] Support of life Main article: Fine-tuned universe The fine-tuned universe hypothesis is the proposition that the conditions that allow the existence of observable life in the universe can only occur when certain universal fundamental physical constants lie within a very narrow range of values. According to this hypothesis, if any of several fundamental constants were only slightly different, the universe would have been unlikely to be conducive to the establishment and development of matter, astronomical structures, elemental diversity, or life as it is understood. Whether this is true, and whether that question is even logically meaningful to ask, are subjects of much debate.[89] The proposition is discussed among philosophers, scientists, theologians, and proponents of creationism.[90] Composition See also: Galaxy formation and evolution, Galaxy cluster, and Nebula The universe is composed almost completely of dark energy, dark matter, and ordinary matter. Other contents are electromagnetic radiation (estimated to constitute from 0.005% to close to 0.01% of the total mass–energy of the universe) and antimatter.[91][92][93] The proportions of all types of matter and energy have changed over the history of the universe.[94] The total amount of electromagnetic radiation generated within the universe has decreased by 1/2 in the past 2 billion years.[95][96] Today, ordinary matter, which includes atoms, stars, galaxies, and life, accounts for only 4.9% of the contents of the universe.[8] The present overall density of this type of matter is very low, roughly 4.5 × 10−31 grams per cubic centimeter, corresponding to a density of the order of only one proton for every four cubic metres of volume.[6] The nature of both dark energy and dark matter is unknown. Dark matter, a mysterious form of matter that has not yet been identified, accounts for 26.8% of the cosmic contents. Dark energy, which is the energy of empty space and is causing the expansion of the universe to accelerate, accounts for the remaining 68.3% of the contents.[8][97][98] The formation of clusters and large-scale filaments in the cold dark matter model with dark energy. The frames show the evolution of structures in a 43 million parsecs (or 140 million light-years) box from redshift of 30 to the present epoch (upper left z=30 to lower right z=0). A map of the superclusters and voids nearest to Earth Matter, dark matter, and dark energy are distributed homogeneously throughout the universe over length scales longer than 300 million light-years (ly) or so.[99] However, over shorter length-scales, matter tends to clump hierarchically; many atoms are condensed into stars, most stars into galaxies, most galaxies into clusters, superclusters and, finally, large-scale galactic filaments. The observable universe contains as many as an estimated 2 trillion galaxies[100][101][102] and, overall, as many as an estimated 1024 stars[103][104] – more stars (and earth-like planets) than all the grains of beach sand on planet Earth;[105][106][107] but less than the total number of atoms estimated in the universe as 1082;[108] and the estimated total number of stars in an inflationary universe (observed and unobserved), as 10100.[109] Typical galaxies range from dwarfs with as few as ten million[110] (107) stars up to giants with one trillion[111] (1012) stars. Between the larger structures are voids, which are typically 10–150 Mpc (33 million–490 million ly) in diameter. The Milky Way is in the Local Group of galaxies, which in turn is in the Laniakea Supercluster.[112] This supercluster spans over 500 million light-years, while the Local Group spans over 10 million light-years.[113] The universe also has vast regions of relative emptiness; the largest known void measures 1.8 billion ly (550 Mpc) across.[114] Comparison of the contents of the universe today to 380,000 years after the Big Bang as measured with 5 year WMAP data (from 2008).[115] Due to rounding errors, the sum of these numbers is not 100%. This reflects the 2008 limits of WMAP's ability to define dark matter and dark energy. The observable universe is isotropic on scales significantly larger than superclusters, meaning that the statistical properties of the universe are the same in all directions as observed from Earth. The universe is bathed in highly isotropic microwave radiation that corresponds to a thermal equilibrium blackbody spectrum of roughly 2.72548 kelvins.[7] The hypothesis that the large-scale universe is homogeneous and isotropic is known as the cosmological principle.[116] A universe that is both homogeneous and isotropic looks the same from all vantage points[117] and has no center.[118] Dark energy Main article: Dark energy An explanation for why the expansion of the universe is accelerating remains elusive. It is often attributed to "dark energy", an unknown form of energy that is hypothesized to permeate space.[119] On a mass–energy equivalence basis, the density of dark energy (~ 7 × 10−30 g/cm3) is much less than the density of ordinary matter or dark matter within galaxies. However, in the present dark-energy era, it dominates the mass–energy of the universe because it is uniform across space.[120][121] Two proposed forms for dark energy are the cosmological constant, a constant energy density filling space homogeneously,[122] and scalar fields such as quintessence or moduli, dynamic quantities whose energy density can vary in time and space while still permeating then enough to cause the observed rate of expansion. Contributions from scalar fields that are constant in space are usually also included in the cosmological constant. The cosmological constant can be formulated to be equivalent to vacuum energy. Scalar fields having only a slight amount of spatial inhomogeneity would be difficult to distinguish from a cosmological constant. Dark matter Main article: Dark matter Dark matter is a hypothetical kind of matter that is invisible to the entire electromagnetic spectrum, but which accounts for most of the matter in the universe. The existence and properties of dark matter are inferred from its gravitational effects on visible matter, radiation, and the large-scale structure of the universe. Other than neutrinos, a form of hot dark matter, dark matter has not been detected directly, making it one of the greatest mysteries in modern astrophysics. Dark matter neither emits nor absorbs light or any other electromagnetic radiation at any significant level. Dark matter is estimated to constitute 26.8% of the total mass–energy and 84.5% of the total matter in the universe.[97][123] Ordinary matter Main article: Matter The remaining 4.9% of the mass–energy of the universe is ordinary matter, that is, atoms, ions, electrons and the objects they form. This matter includes stars, which produce nearly all of the light we see from galaxies, as well as interstellar gas in the interstellar and intergalactic media, planets, and all the objects from everyday life that we can bump into, touch or squeeze.[124] The great majority of ordinary matter in the universe is unseen, since visible stars and gas inside galaxies and clusters account for less than 10 percent of the ordinary matter contribution to the mass–energy density of the universe.[125][126][127] Ordinary matter commonly exists in four states (or phases): solid, liquid, gas, and plasma.[128] However, advances in experimental techniques have revealed other previously theoretical phases, such as Bose–Einstein condensates and fermionic condensates.[129][130] Ordinary matter is composed of two types of elementary particles: quarks and leptons.[131] For example, the proton is formed of two up quarks and one down quark; the neutron is formed of two down quarks and one up quark; and the electron is a kind of lepton. An atom consists of an atomic nucleus, made up of protons and neutrons (both of which are baryons), and electrons that orbit the nucleus.[46]: 1476  Because most of the mass of an atom is concentrated in its nucleus, which is made up of baryons, astronomers often use the term baryonic matter to describe ordinary matter, although a small fraction of this "baryonic matter" is electrons. Soon after the Big Bang, primordial protons and neutrons formed from the quark–gluon plasma of the early universe as it cooled below two trillion degrees. A few minutes later, in a process known as Big Bang nucleosynthesis, nuclei formed from the primordial protons and neutrons. This nucleosynthesis formed lighter elements, those with small atomic numbers up to lithium and beryllium, but the abundance of heavier elements dropped off sharply with increasing atomic number. Some boron may have been formed at this time, but the next heavier element, carbon, was not formed in significant amounts. Big Bang nucleosynthesis shut down after about 20 minutes due to the rapid drop in temperature and density of the expanding universe. Subsequent formation of heavier elements resulted from stellar nucleosynthesis and supernova nucleosynthesis.[132] Particles A four-by-four table of particles. Columns are three generations of matter (fermions) and one of forces (bosons). In the first three columns, two rows contain quarks and two leptons. The top two rows' columns contain up (u) and down (d) quarks, charm (c) and strange (s) quarks, top (t) and bottom (b) quarks, and photon (γ) and gluon (g), respectively. The bottom two rows' columns contain electron neutrino (ν sub e) and electron (e), muon neutrino (ν sub μ) and muon (μ), and tau neutrino (ν sub τ) and tau (τ), and Z sup 0 and W sup ± weak force. Mass, charge, and spin are listed for each particle. Standard model of elementary particles: the 12 fundamental fermions and 4 fundamental bosons. Brown loops indicate which bosons (red) couple to which fermions (purple and green). Columns are three generations of matter (fermions) and one of forces (bosons). In the first three columns, two rows contain quarks and two leptons. The top two rows' columns contain up (u) and down (d) quarks, charm (c) and strange (s) quarks, top (t) and bottom (b) quarks, and photon (γ) and gluon (g), respectively. The bottom two rows' columns contain electron neutrino (νe) and electron (e), muon neutrino (νμ) and muon (μ), tau neutrino (ντ) and tau (τ), and the Z0 and W± carriers of the weak force. Mass, charge, and spin are listed for each particle. Main article: Particle physics Ordinary matter and the forces that act on matter can be described in terms of elementary particles.[133] These particles are sometimes described as being fundamental, since they have an unknown substructure, and it is unknown whether or not they are composed of smaller and even more fundamental particles.[134][135] In most contemporary models they are thought of as points in space.[136] All elementary particles are currently best explained by quantum mechanics and exhibit wave–particle duality: their behavior has both particle-like and wave-like aspects, with different features dominating under different circumstances.[137] Of central importance is the Standard Model, a theory that is concerned with electromagnetic interactions and the weak and strong nuclear interactions.[138] The Standard Model is supported by the experimental confirmation of the existence of particles that compose matter: quarks and leptons, and their corresponding "antimatter" duals, as well as the force particles that mediate interactions: the photon, the W and Z bosons, and the gluon.[134] The Standard Model predicted the existence of the recently discovered Higgs boson, a particle that is a manifestation of a field within the universe that can endow particles with mass.[139][140] Because of its success in explaining a wide variety of experimental results, the Standard Model is sometimes regarded as a "theory of almost everything".[138] The Standard Model does not, however, accommodate gravity. A true force–particle "theory of everything" has not been attained.[141] Hadrons Main article: Hadron A hadron is a composite particle made of quarks held together by the strong force. Hadrons are categorized into two families: baryons (such as protons and neutrons) made of three quarks, and mesons (such as pions) made of one quark and one antiquark. Of the hadrons, protons are stable, and neutrons bound within atomic nuclei are stable. Other hadrons are unstable under ordinary conditions and are thus insignificant constituents of the modern universe.[142]: 118–123  From approximately 10−6 seconds after the Big Bang, during a period known as the hadron epoch, the temperature of the universe had fallen sufficiently to allow quarks to bind together into hadrons, and the mass of the universe was dominated by hadrons. Initially, the temperature was high enough to allow the formation of hadron–anti-hadron pairs, which kept matter and antimatter in thermal equilibrium. However, as the temperature of the universe continued to fall, hadron–anti-hadron pairs were no longer produced. Most of the hadrons and anti-hadrons were then eliminated in particle–antiparticle annihilation reactions, leaving a small residual of hadrons by the time the universe was about one second old.[142]: 244–266  Leptons Main article: Lepton A lepton is an elementary, half-integer spin particle that does not undergo strong interactions but is subject to the Pauli exclusion principle; no two leptons of the same species can be in exactly the same state at the same time.[143] Two main classes of leptons exist: charged leptons (also known as the electron-like leptons), and neutral leptons (better known as neutrinos). Electrons are stable and the most common charged lepton in the universe, whereas muons and taus are unstable particles that quickly decay after being produced in high energy collisions, such as those involving cosmic rays or carried out in particle accelerators.[144][145] Charged leptons can combine with other particles to form various composite particles such as atoms and positronium. The electron governs nearly all of chemistry, as it is found in atoms and is directly tied to all chemical properties. Neutrinos rarely interact with anything, and are consequently rarely observed. Neutrinos stream throughout the universe but rarely interact with normal matter.[146] The lepton epoch was the period in the evolution of the early universe in which the leptons dominated the mass of the universe. It started roughly 1 second after the Big Bang, after the majority of hadrons and anti-hadrons annihilated each other at the end of the hadron epoch. During the lepton epoch the temperature of the universe was still high enough to create lepton–anti-lepton pairs, so leptons and anti-leptons were in thermal equilibrium. Approximately 10 seconds after the Big Bang, the temperature of the universe had fallen to the point where lepton–anti-lepton pairs were no longer created.[147] Most leptons and anti-leptons were then eliminated in annihilation reactions, leaving a small residue of leptons. The mass of the universe was then dominated by photons as it entered the following photon epoch.[148][149] Photons Main article: Photon epoch See also: Photino A photon is the quantum of light and all other forms of electromagnetic radiation. It is the carrier for the electromagnetic force. The effects of this force are easily observable at the microscopic and at the macroscopic level because the photon has zero rest mass; this allows long distance interactions.[46]: 1470  The photon epoch started after most leptons and anti-leptons were annihilated at the end of the lepton epoch, about 10 seconds after the Big Bang. Atomic nuclei were created in the process of nucleosynthesis which occurred during the first few minutes of the photon epoch. For the remainder of the photon epoch the universe contained a hot dense plasma of nuclei, electrons and photons. About 380,000 years after the Big Bang, the temperature of the universe fell to the point where nuclei could combine with electrons to create neutral atoms. As a result, photons no longer interacted frequently with matter and the universe became transparent. The highly redshifted photons from this period form the cosmic microwave background. Tiny variations in temperature and density detectable in the CMB were the early "seeds" from which all subsequent structure formation took place.[142]: 244–266  vte Timeline of the Big Bang Cosmological models Model of the universe based on general relativity Main article: Solutions of the Einstein field equations See also: Big Bang and Ultimate fate of the universe General relativity is the geometric theory of gravitation published by Albert Einstein in 1915 and the current description of gravitation in modern physics. It is the basis of current cosmological models of the universe. General relativity generalizes special relativity and Newton's law of universal gravitation, providing a unified description of gravity as a geometric property of space and time, or spacetime. In particular, the curvature of spacetime is directly related to the energy and momentum of whatever matter and radiation are present.[150] The relation is specified by the Einstein field equations, a system of partial differential equations. In general relativity, the distribution of matter and energy determines the geometry of spacetime, which in turn describes the acceleration of matter. Therefore, solutions of the Einstein field equations describe the evolution of the universe. Combined with measurements of the amount, type, and distribution of matter in the universe, the equations of general relativity describe the evolution of the universe over time.[150] With the assumption of the cosmological principle that the universe is homogeneous and isotropic everywhere, a specific solution of the field equations that describes the universe is the metric tensor called the Friedmann–Lemaître–Robertson–Walker metric, d s 2 = − c 2 d t 2 + R ( t ) 2 ( d r 2 1 − k r 2 + r 2 d θ 2 + r 2 sin 2 ⁡ θ d ϕ 2 ) {\displaystyle ds^{2}=-c^{2}dt^{2}+R(t)^{2}\left({\frac {dr^{2}}{1-kr^{2}}}+r^{2}d\theta ^{2}+r^{2}\sin ^{2}\theta \,d\phi ^{2}\right)} where (r, θ, φ) correspond to a spherical coordinate system. This metric has only two undetermined parameters. An overall dimensionless length scale factor R describes the size scale of the universe as a function of time (an increase in R is the expansion of the universe),[151] and a curvature index k describes the geometry. The index k is defined so that it can take only one of three values: 0, corresponding to flat Euclidean geometry; 1, corresponding to a space of positive curvature; or −1, corresponding to a space of positive or negative curvature.[152] The value of R as a function of time t depends upon k and the cosmological constant Λ.[150] The cosmological constant represents the energy density of the vacuum of space and could be related to dark energy.[98] The equation describing how R varies with time is known as the Friedmann equation after its inventor, Alexander Friedmann.[153] The solutions for R(t) depend on k and Λ, but some qualitative features of such solutions are general. First and most importantly, the length scale R of the universe can remain constant only if the universe is perfectly isotropic with positive curvature (k=1) and has one precise value of density everywhere, as first noted by Albert Einstein.[150] However, this equilibrium is unstable: if the density were slightly different from the needed value, at any place, the difference would be amplified over time. Second, all solutions suggest that there was a gravitational singularity in the past, when R went to zero and matter and energy were infinitely dense. It may seem that this conclusion is uncertain because it is based on the questionable assumptions of perfect homogeneity and isotropy (the cosmological principle) and that only the gravitational interaction is significant. However, the Penrose–Hawking singularity theorems show that a singularity should exist for very general conditions. Hence, according to Einstein's field equations, R grew rapidly from an unimaginably hot, dense state that existed immediately following this singularity (when R had a small, finite value); this is the essence of the Big Bang model of the universe. Understanding the singularity of the Big Bang likely requires a quantum theory of gravity, which has not yet been formulated.[154] Third, the curvature index k determines the sign of the curvature of constant-time spatial surfaces[152] averaged over sufficiently large length scales (greater than about a billion light-years). If k=1, the curvature is positive and the universe has a finite volume.[155] A universe with positive curvature is often visualized as a three-dimensional sphere embedded in a four-dimensional space. Conversely, if k is zero or negative, the universe has an infinite volume.[155] It may seem counter-intuitive that an infinite and yet infinitely dense universe could be created in a single instant when R = 0, but exactly that is predicted mathematically when k is nonpositive and the cosmological principle is satisfied. By analogy, an infinite plane has zero curvature but infinite area, whereas an infinite cylinder is finite in one direction and a torus is finite in both. A toroidal universe could behave like a normal universe with periodic boundary conditions. The ultimate fate of the universe is still unknown because it depends critically on the curvature index k and the cosmological constant Λ. If the universe were sufficiently dense, k would equal +1, meaning that its average curvature throughout is positive and the universe will eventually recollapse in a Big Crunch,[156] possibly starting a new universe in a Big Bounce. Conversely, if the universe were insufficiently dense, k would equal 0 or −1 and the universe would expand forever, cooling off and eventually reaching the Big Freeze and the heat death of the universe.[150] Modern data suggests that the expansion of the universe is accelerating; if this acceleration is sufficiently rapid, the universe may eventually reach a Big Rip. Observationally, the universe appears to be flat (k = 0), with an overall density that is very close to the critical value between recollapse and eternal expansion.[157] Multiverse hypotheses Main articles: Multiverse, Many-worlds interpretation, and Bubble universe theory See also: Eternal inflation Some speculative theories have proposed that our universe is but one of a set of disconnected universes, collectively denoted as the multiverse, challenging or enhancing more limited definitions of the universe.[19][158] Scientific multiverse models are distinct from concepts such as alternate planes of consciousness and simulated reality. Max Tegmark developed a four-part classification scheme for the different types of multiverses that scientists have suggested in response to various problems in physics. An example of such multiverses is the one resulting from the chaotic inflation model of the early universe.[159] Another is the multiverse resulting from the many-worlds interpretation of quantum mechanics. In this interpretation, parallel worlds are generated in a manner similar to quantum superposition and decoherence, with all states of the wave functions being realized in separate worlds. Effectively, in the many-worlds interpretation the multiverse evolves as a universal wavefunction. If the Big Bang that created our multiverse created an ensemble of multiverses, the wave function of the ensemble would be entangled in this sense.[160] Whether scientifically meaningful probabilities can be extracted from this picture has been and continues to be a topic of much debate, and multiple versions of the many-worlds interpretation exist.[161][162][163] (The subject of the interpretation of quantum mechanics is in general marked by disagreement.)[164][165][166] The least controversial, but still highly disputed, category of multiverse in Tegmark's scheme is Level I. The multiverses of this level are composed by distant spacetime events "in our own universe". Tegmark and others[167] have argued that, if space is infinite, or sufficiently large and uniform, identical instances of the history of Earth's entire Hubble volume occur every so often, simply by chance. Tegmark calculated that our nearest so-called doppelgänger is 1010115 metres away from us (a double exponential function larger than a googolplex).[168][169] However, the arguments used are of speculative nature.[170] Additionally, it would be impossible to scientifically verify the existence of an identical Hubble volume. It is possible to conceive of disconnected spacetimes, each existing but unable to interact with one another.[168][171] An easily visualized metaphor of this concept is a group of separate soap bubbles, in which observers living on one soap bubble cannot interact with those on other soap bubbles, even in principle.[172] According to one common terminology, each "soap bubble" of spacetime is denoted as a universe, whereas humans' particular spacetime is denoted as the universe,[19] just as humans call Earth's moon the Moon. The entire collection of these separate spacetimes is denoted as the multiverse.[19] With this terminology, different universes are not causally connected to each other.[19] In principle, the other unconnected universes may have different dimensionalities and topologies of spacetime, different forms of matter and energy, and different physical laws and physical constants, although such possibilities are purely speculative.[19] Others consider each of several bubbles created as part of chaotic inflation to be separate universes, though in this model these universes all share a causal origin.[19] Historical conceptions See also: Cosmology, Timeline of cosmological theories, Nicolaus Copernicus § Copernican system, and Philosophiæ Naturalis Principia Mathematica § Beginnings of the Scientific Revolution Historically, there have been many ideas of the cosmos (cosmologies) and its origin (cosmogonies). Theories of an impersonal universe governed by physical laws were first proposed by the Greeks and Indians.[13] Ancient Chinese philosophy encompassed the notion of the universe including both all of space and all of time.[173] Over the centuries, improvements in astronomical observations and theories of motion and gravitation led to ever more accurate descriptions of the universe. The modern era of cosmology began with Albert Einstein's 1915 general theory of relativity, which made it possible to quantitatively predict the origin, evolution, and conclusion of the universe as a whole. Most modern, accepted theories of cosmology are based on general relativity and, more specifically, the predicted Big Bang.[174] Mythologies Main articles: Creation myth, Cosmogony, and Religious cosmology Many cultures have stories describing the origin of the world and universe. Cultures generally regard these stories as having some truth. There are however many differing beliefs in how these stories apply amongst those believing in a supernatural origin, ranging from a god directly creating the universe as it is now to a god just setting the "wheels in motion" (for example via mechanisms such as the big bang and evolution).[175] Ethnologists and anthropologists who study myths have developed various classification schemes for the various themes that appear in creation stories.[176][177] For example, in one type of story, the world is born from a world egg; such stories include the Finnish epic poem Kalevala, the Chinese story of Pangu or the Indian Brahmanda Purana. In related stories, the universe is created by a single entity emanating or producing something by him- or herself, as in the Tibetan Buddhism concept of Adi-Buddha, the ancient Greek story of Gaia (Mother Earth), the Aztec goddess Coatlicue myth, the ancient Egyptian god Atum story, and the Judeo-Christian Genesis creation narrative in which the Abrahamic God created the universe. In another type of story, the universe is created from the union of male and female deities, as in the Maori story of Rangi and Papa. In other stories, the universe is created by crafting it from pre-existing materials, such as the corpse of a dead god—as from Tiamat in the Babylonian epic Enuma Elish or from the giant Ymir in Norse mythology—or from chaotic materials, as in Izanagi and Izanami in Japanese mythology. In other stories, the universe emanates from fundamental principles, such as Brahman and Prakrti, the creation myth of the Serers,[178] or the yin and yang of the Tao. Philosophical models Further information: Cosmology See also: Pre-Socratic philosophy, Physics (Aristotle), Hindu cosmology, Islamic cosmology, and Philosophy of space and time The pre-Socratic Greek philosophers and Indian philosophers developed some of the earliest philosophical concepts of the universe.[13][179] The earliest Greek philosophers noted that appearances can be deceiving, and sought to understand the underlying reality behind the appearances. In particular, they noted the ability of matter to change forms (e.g., ice to water to steam) and several philosophers proposed that all the physical materials in the world are different forms of a single primordial material, or arche. The first to do so was Thales, who proposed this material to be water. Thales' student, Anaximander, proposed that everything came from the limitless apeiron. Anaximenes proposed the primordial material to be air on account of its perceived attractive and repulsive qualities that cause the arche to condense or dissociate into different forms. Anaxagoras proposed the principle of Nous (Mind), while Heraclitus proposed fire (and spoke of logos). Empedocles proposed the elements to be earth, water, air and fire. His four-element model became very popular. Like Pythagoras, Plato believed that all things were composed of number, with Empedocles' elements taking the form of the Platonic solids. Democritus, and later philosophers—most notably Leucippus—proposed that the universe is composed of indivisible atoms moving through a void (vacuum), although Aristotle did not believe that to be feasible because air, like water, offers resistance to motion. Air will immediately rush in to fill a void, and moreover, without resistance, it would do so indefinitely fast.[13] Although Heraclitus argued for eternal change,[180] his contemporary Parmenides emphasized changelessness. Parmenides' poem On Nature has been read as saying that all change is an illusion, that the true underlying reality is eternally unchanging and of a single nature, or at least that the essential feature of each thing that exists must exist eternally, without origin, change, or end.[181] His student Zeno of Elea challenged everyday ideas about motion with several famous paradoxes. Aristotle responded to these paradoxes by developing the notion of a potential countable infinity, as well as the infinitely divisible continuum.[182][183] Unlike the eternal and unchanging cycles of time, he believed that the world is bounded by the celestial spheres and that cumulative stellar magnitude is only finitely multiplicative. The Indian philosopher Kanada, founder of the Vaisheshika school, developed a notion of atomism and proposed that light and heat were varieties of the same substance.[184] In the 5th century AD, the Buddhist atomist philosopher Dignāga proposed atoms to be point-sized, durationless, and made of energy. They denied the existence of substantial matter and proposed that movement consisted of momentary flashes of a stream of energy.[185] The notion of temporal finitism was inspired by the doctrine of creation shared by the three Abrahamic religions: Judaism, Christianity and Islam. The Christian philosopher, John Philoponus, presented the philosophical arguments against the ancient Greek notion of an infinite past and future. Philoponus' arguments against an infinite past were used by the early Muslim philosopher, Al-Kindi (Alkindus); the Jewish philosopher, Saadia Gaon (Saadia ben Joseph); and the Muslim theologian, Al-Ghazali (Algazel).[186] Pantheism is the philosophical religious belief that the universe itself is identical to divinity and a supreme being or entity.[187] The physical universe is thus understood as an all-encompassing, immanent deity.[188] The term 'pantheist' designates one who holds both that everything constitutes a unity and that this unity is divine, consisting of an all-encompassing, manifested god or goddess.[189][190] Pantheistic concepts date back thousands of years, and pantheistic elements have been identified in various religious traditions. Astronomical concepts Main articles: History of astronomy and Timeline of astronomy 3rd century BCE calculations by Aristarchus on the relative sizes of, from left to right, the Sun, Earth, and Moon, from a 10th-century AD Greek copy The earliest written records of identifiable predecessors to modern astronomy come from Ancient Egypt and Mesopotamia from around 3000 to 1200 BCE.[191][192] Babylonian astronomers of the 7th century BCE viewed the world as a flat disk surrounded by the ocean,[193][194] and this forms the premise for early Greek maps like those of Anaximander and Hecataeus of Miletus. Later Greek philosophers, observing the motions of the heavenly bodies, were concerned with developing models of the universe based more profoundly on empirical evidence. The first coherent model was proposed by Eudoxus of Cnidos, a student of Plato who followed Plato's idea that heavenly motions had to be circular. In order to account for the known complications of the planets' motions, particularly retrograde movement, Eudoxus' model included 27 different celestial spheres: four for each of the planets visible to the naked eye, three each for the Sun and the Moon, and one for the stars. All of these spheres were centered on the Earth, which remained motionless while they rotated eternally. Aristotle elaborated upon this model, increasing the number of spheres to 55 in order to account for further details of planetary motion. For Aristotle, normal matter was entirely contained within the terrestrial sphere, and it obeyed fundamentally different rules from heavenly material.[195][196] The post-Aristotle treatise De Mundo (of uncertain authorship and date) stated, "Five elements, situated in spheres in five regions, the less being in each case surrounded by the greater—namely, earth surrounded by water, water by air, air by fire, and fire by ether—make up the whole universe".[197] This model was also refined by Callippus and after concentric spheres were abandoned, it was brought into nearly perfect agreement with astronomical observations by Ptolemy.[198] The success of such a model is largely due to the mathematical fact that any function (such as the position of a planet) can be decomposed into a set of circular functions (the Fourier modes). Other Greek scientists, such as the Pythagorean philosopher Philolaus, postulated (according to Stobaeus' account) that at the center of the universe was a "central fire" around which the Earth, Sun, Moon and planets revolved in uniform circular motion.[199] The Greek astronomer Aristarchus of Samos was the first known individual to propose a heliocentric model of the universe. Though the original text has been lost, a reference in Archimedes' book The Sand Reckoner describes Aristarchus's heliocentric model. Archimedes wrote: You, King Gelon, are aware the universe is the name given by most astronomers to the sphere the center of which is the center of the Earth, while its radius is equal to the straight line between the center of the Sun and the center of the Earth. This is the common account as you have heard from astronomers. But Aristarchus has brought out a book consisting of certain hypotheses, wherein it appears, as a consequence of the assumptions made, that the universe is many times greater than the universe just mentioned. His hypotheses are that the fixed stars and the Sun remain unmoved, that the Earth revolves about the Sun on the circumference of a circle, the Sun lying in the middle of the orbit, and that the sphere of fixed stars, situated about the same center as the Sun, is so great that the circle in which he supposes the Earth to revolve bears such a proportion to the distance of the fixed stars as the center of the sphere bears to its surface.[200] Aristarchus thus believed the stars to be very far away, and saw this as the reason why stellar parallax had not been observed, that is, the stars had not been observed to move relative each other as the Earth moved around the Sun. The stars are in fact much farther away than the distance that was generally assumed in ancient times, which is why stellar parallax is only detectable with precision instruments. The geocentric model, consistent with planetary parallax, was assumed to be the explanation for the unobservability of stellar parallax.[201] Flammarion engraving, Paris 1888 The only other astronomer from antiquity known by name who supported Aristarchus's heliocentric model was Seleucus of Seleucia, a Hellenistic astronomer who lived a century after Aristarchus.[202][203][204] According to Plutarch, Seleucus was the first to prove the heliocentric system through reasoning, but it is not known what arguments he used. Seleucus' arguments for a heliocentric cosmology were probably related to the phenomenon of tides.[205] According to Strabo (1.1.9), Seleucus was the first to state that the tides are due to the attraction of the Moon, and that the height of the tides depends on the Moon's position relative to the Sun.[206] Alternatively, he may have proved heliocentricity by determining the constants of a geometric model for it, and by developing methods to compute planetary positions using this model, similar to Nicolaus Copernicus in the 16th century.[207] During the Middle Ages, heliocentric models were also proposed by the Persian astronomers Albumasar[208] and Al-Sijzi.[209] Model of the Copernican Universe by Thomas Digges in 1576, with the amendment that the stars are no longer confined to a sphere, but spread uniformly throughout the space surrounding the planets The Aristotelian model was accepted in the Western world for roughly two millennia, until Copernicus revived Aristarchus's perspective that the astronomical data could be explained more plausibly if the Earth rotated on its axis and if the Sun were placed at the center of the universe.[210] In the center rests the Sun. For who would place this lamp of a very beautiful temple in another or better place than this wherefrom it can illuminate everything at the same time? — Nicolaus Copernicus, in Chapter 10, Book 1 of De Revolutionibus Orbium Coelestrum (1543) As noted by Copernicus, the notion that the Earth rotates is very old, dating at least to Philolaus (c. 450 BC), Heraclides Ponticus (c. 350 BC) and Ecphantus the Pythagorean. Roughly a century before Copernicus, the Christian scholar Nicholas of Cusa also proposed that the Earth rotates on its axis in his book, On Learned Ignorance (1440).[211] Al-Sijzi[212] also proposed that the Earth rotates on its axis. Empirical evidence for the Earth's rotation on its axis, using the phenomenon of comets, was given by Tusi (1201–1274) and Ali Qushji (1403–1474).[213] This cosmology was accepted by Isaac Newton, Christiaan Huygens and later scientists.[214] Newton demonstrated that the same laws of motion and gravity apply to earthly and to celestial matter, making Aristotle's division between the two obsolete. Edmund Halley (1720)[215] and Jean-Philippe de Chéseaux (1744)[216] noted independently that the assumption of an infinite space filled uniformly with stars would lead to the prediction that the nighttime sky would be as bright as the Sun itself; this became known as Olbers' paradox in the 19th century.[217] Newton believed that an infinite space uniformly filled with matter would cause infinite forces and instabilities causing the matter to be crushed inwards under its own gravity.[214] This instability was clarified in 1902 by the Jeans instability criterion.[218] One solution to these paradoxes is the Charlier universe, in which the matter is arranged hierarchically (systems of orbiting bodies that are themselves orbiting in a larger system, ad infinitum) in a fractal way such that the universe has a negligibly small overall density; such a cosmological model had also been proposed earlier in 1761 by Johann Heinrich Lambert.[53][219] During the 18th century, Immanuel Kant speculated that nebulae could be entire galaxies separate from the Milky Way,[215] and in 1850, Alexander von Humboldt called these separate galaxies Weltinseln, or "world islands", a term that later developed into "island universes".[220][221] In 1919, when the Hooker Telescope was completed, the prevailing view was that the universe consisted entirely of the Milky Way Galaxy. Using the Hooker Telescope, Edwin Hubble identified Cepheid variables in several spiral nebulae and in 1922–1923 proved conclusively that Andromeda Nebula and Triangulum among others, were entire galaxies outside our own, thus proving that the universe consists of a multitude of galaxies.[222] The modern era of physical cosmology began in 1917, when Albert Einstein first applied his general theory of relativity to model the structure and dynamics of the universe.[223] The discoveries of this era, and the questions that remain unanswered, are outlined in the sections above. Map of the observable universe with some of the notable astronomical objects known as of 2018. The scale of length increases exponentially toward the right. Celestial bodies are shown enlarged in size to be able to understand their shapes. Location of the Earth in the universe Earth Solar System Radcliffe Wave Orion Arm Milky Way Local Group Virgo SCl Laniakea SCl Observable universe See also Cosmic Calendar (scaled down timeline) Cosmic latte Detailed logarithmic timeline Earth's location in the universe False vacuum Future of an expanding universe Galaxy And Mass Assembly survey Heat death of the universe History of the center of the Universe Illustris project Non-standard cosmology Nucleocosmochronology Parallel universe (fiction) Rare Earth hypothesis Space and survival Terasecond and longer Timeline of the early universe Timeline of the far future Timeline of the near future Zero-energy universe References Footnotes According to modern physics, particularly the theory of relativity, space and time are intrinsically linked as spacetime. Although listed in megaparsecs by the cited source, this number is so vast that its digits would remain virtually unchanged for all intents and purposes regardless of which conventional units it is listed in, whether it to be nanometers or gigaparsecs, as the differences would disappear into the error. Citations "Hubble sees galaxies galore". spacetelescope.org. Archived from the original on May 4, 2017. Retrieved April 30, 2017. Planck Collaboration (2016). "Planck 2015 results. XIII. Cosmological parameters". Astronomy & Astrophysics. 594: A13, Table 4. arXiv:1502.01589. Bibcode:2016A&A...594A..13P. doi:10.1051/0004-6361/201525830. S2CID 119262962. Greene, Brian (2011). The Hidden Reality. Alfred A. Knopf. Bars, Itzhak; Terning, John (2009). Extra Dimensions in Space and Time. Springer. pp. 27–. ISBN 978-0-387-77637-8. Retrieved May 1, 2011. Davies, Paul (2006). The Goldilocks Enigma. First Mariner Books. pp. 43ff. ISBN 978-0-618-59226-5. NASA/WMAP Science Team (January 24, 2014). "Universe 101: What is the Universe Made Of?". NASA. Archived from the original on March 10, 2008. Retrieved February 17, 2015. Fixsen, D.J. (2009). "The Temperature of the Cosmic Microwave Background". The Astrophysical Journal. 707 (2): 916–920. arXiv:0911.1955. Bibcode:2009ApJ...707..916F. doi:10.1088/0004-637X/707/2/916. S2CID 119217397. "First Planck results: the universe is still weird and interesting". Matthew Francis. Ars technica. March 21, 2013. Archived from the original on May 2, 2019. Retrieved August 21, 2015. NASA/WMAP Science Team (January 24, 2014). "Universe 101: Will the Universe expand forever?". NASA. Archived from the original on March 9, 2008. Retrieved April 16, 2015. Zeilik, Michael; Gregory, Stephen A. (1998). Introductory Astronomy & Astrophysics (4th ed.). Saunders College Publishing. ISBN 978-0-03-006228-5. "The totality of all space and time; all that is, has been, and will be." Planck Collaboration; Aghanim, N.; Akrami, Y.; Ashdown, M.; Aumont, J.; Baccigalupi, C.; Ballardini, M.; Banday, A. J.; Barreiro, R. B.; Bartolo, N.; Basak, S. (September 2020). "Planck 2018 results: VI. Cosmological parameters". Astronomy & Astrophysics. 641: A6. arXiv:1807.06209. Bibcode:2020A&A...641A...6P. doi:10.1051/0004-6361/201833910. ISSN 0004-6361. S2CID 119335614. Dold-Samplonius, Yvonne (2002). From China to Paris: 2000 Years Transmission of Mathematical Ideas. Franz Steiner Verlag. Glick, Thomas F.; Livesey, Steven; Wallis, Faith (2005). Medieval Science Technology and Medicine: An Encyclopedia. Routledge. ISBN 978-0-415-96930-7. OCLC 61228669. Carroll, Bradley W.; Ostlie, Dale A. (2013). An Introduction to Modern Astrophysics (International ed.). Pearson. pp. 1173–1174. ISBN 978-1-292-02293-2. Archived from the original on December 28, 2019. Retrieved May 16, 2018. Hawking, Stephen (1988). A Brief History of Time. Bantam Books. p. 43. ISBN 978-0-553-05340-1. Redd, Nola. "What is Dark Matter?". Space.com. Archived from the original on February 1, 2018. Retrieved February 1, 2018. "Planck 2015 results, table 9". Archived from the original on July 27, 2018. Retrieved May 16, 2018. Persic, Massimo; Salucci, Paolo (September 1, 1992). "The baryon content of the Universe". Monthly Notices of the Royal Astronomical Society. 258 (1): 14P–18P. arXiv:astro-ph/0502178. Bibcode:1992MNRAS.258P..14P. doi:10.1093/mnras/258.1.14P. ISSN 0035-8711. S2CID 17945298. Ellis, George F. R.; Kirchner, U.; Stoeger, W. R. (2004). "Multiverses and physical cosmology". Monthly Notices of the Royal Astronomical Society. 347 (3): 921–936. arXiv:astro-ph/0305292. Bibcode:2004MNRAS.347..921E. doi:10.1111/j.1365-2966.2004.07261.x. S2CID 119028830. "'Multiverse' theory suggested by microwave background". BBC News. August 3, 2011. Archived from the original on February 14, 2023. Retrieved February 14, 2023. "Universe". Encyclopaedia Britannica online. 2012. Archived from the original on June 9, 2021. Retrieved February 17, 2018. "Universe". Merriam-Webster Dictionary. Archived from the original on October 22, 2012. Retrieved September 21, 2012. "Universe". Dictionary.com. Archived from the original on October 23, 2012. Retrieved September 21, 2012. Schreuder, Duco A. (2014). Vision and Visual Perception. Archway Publishing. p. 135. ISBN 978-1-4808-1294-9. Archived from the original on April 22, 2021. Retrieved January 27, 2016. Mermin, N. David (2004). "Could Feynman Have Said This?". Physics Today. 57 (5): 10. Bibcode:2004PhT....57e..10M. doi:10.1063/1.1768652. Tegmark, Max (2008). "The Mathematical Universe". Foundations of Physics. 38 (2): 101–150. arXiv:0704.0646. Bibcode:2008FoPh...38..101T. doi:10.1007/s10701-007-9186-9. S2CID 9890455. A short version of which is available at Fixsen, D. J. (2007). "Shut up and calculate". arXiv:0709.4024 [physics.pop-ph]. in reference to David Mermin's famous quote "shut up and calculate!"[25] Holt, Jim (2012). Why Does the World Exist?. Liveright Publishing. p. 308. Ferris, Timothy (1997). The Whole Shebang: A State-of-the-Universe(s) Report. Simon & Schuster. p. 400. Copan, Paul; William Lane Craig (2004). Creation Out of Nothing: A Biblical, Philosophical, and Scientific Exploration. Baker Academic. p. 220. ISBN 978-0-8010-2733-8. Bolonkin, Alexander (2011). Universe, Human Immortality and Future Human Evaluation. Elsevier. pp. 3–. ISBN 978-0-12-415801-6. Archived from the original on February 8, 2021. Retrieved January 27, 2016. The Compact Edition of the Oxford English Dictionary, volume II, Oxford: Oxford University Press, 1971, p. 3518. ISBN 978-0198611172. Lewis, C.T. and Short, S (1879) A Latin Dictionary, Oxford University Press, ISBN 0-19-864201-6, pp. 1933, 1977–1978. Liddell; Scott. "A Greek-English Lexicon". lsj.gr. Archived from the original on November 6, 2018. Retrieved July 30, 2022. "πᾶς" Liddell; Scott. "A Greek-English Lexicon". lsj.gr. Archived from the original on November 6, 2018. Retrieved July 30, 2022. "ὅλος" Liddell; Scott. "A Greek–English Lexicon". lsj.gr. Archived from the original on November 6, 2018. Retrieved July 30, 2022. "κόσμος" Lewis, C.T.; Short, S (1879). A Latin Dictionary. Oxford University Press. pp. 1175, 1189–1190, 1881–1882. ISBN 978-0-19-864201-5. The Compact Edition of the Oxford English Dictionary. Vol. II. Oxford: Oxford University Press. 1971. pp. 569, 909, 1900, 3821–3822. ISBN 978-0-19-861117-2. Silk, Joseph (2009). Horizons of Cosmology. Templeton Pressr. p. 208. Singh, Simon (2005). Big Bang: The Origin of the Universe. Harper Perennial. p. 560. Bibcode:2004biba.book.....S. Sivaram, C. (1986). "Evolution of the Universe through the Planck epoch". Astrophysics and Space Science. 125 (1): 189–199. Bibcode:1986Ap&SS.125..189S. doi:10.1007/BF00643984. S2CID 123344693. Johnson, Jennifer A. (February 2019). "Populating the periodic table: Nucleosynthesis of the elements". Science. 363 (6426): 474–478. Bibcode:2019Sci...363..474J. doi:10.1126/science.aau9540. ISSN 0036-8075. PMID 30705182. S2CID 59565697. Durrer, Ruth (2008). The Cosmic Microwave Background. Cambridge University Press. ISBN 978-0-521-84704-9. Steane, Andrew M. (2021). Relativity Made Relatively Easy, Volume 2: General Relativity and Cosmology. Oxford University Press. ISBN 978-0-192-89564-6. Larson, Richard B. & Bromm, Volker (March 2002). "The First Stars in the Universe". Scientific American. Archived from the original on June 11, 2015. Retrieved June 9, 2015. Ryden, Barbara, "Introduction to Cosmology", 2006, eqn. 6.33 Urone, Paul Peter; et al. (2022). College Physics 2e. OpenStax. ISBN 978-1-951-69360-2. Archived from the original on February 13, 2023. Retrieved February 13, 2023. "Antimatter". Particle Physics and Astronomy Research Council. October 28, 2003. Archived from the original on March 7, 2004. Retrieved August 10, 2006. Smorra C.; et al. (October 20, 2017). "A parts-per-billion measurement of the antiproton magnetic moment" (PDF). Nature. 550 (7676): 371–374. Bibcode:2017Natur.550..371S. doi:10.1038/nature24048. PMID 29052625. S2CID 205260736. Archived (PDF) from the original on October 30, 2018. Retrieved August 25, 2019. Landau & Lifshitz (1975, p. 361): "It is interesting to note that in a closed space the total electric charge must be zero. Namely, every closed surface in a finite space encloses on each side of itself a finite region of space. Therefore, the flux of the electric field through this surface is equal, on the one hand, to the total charge located in the interior of the surface, and on the other hand to the total charge outside of it, with opposite sign. Consequently, the sum of the charges on the two sides of the surface is zero." Kaku, Michio (2008). Physics of the Impossible: A Scientific Exploration into the World of Phasers, Force Fields, Teleportation, and Time Travel. Knopf Doubleday Publishing Group. pp. 202–. ISBN 978-0-385-52544-2. Bars, Itzhak; Terning, John (2018). Extra Dimensions in Space and Time. Springer. pp. 27–. ISBN 978-0-387-77637-8. Retrieved October 19, 2018. Crockett, Christopher (February 20, 2013). "What is a light-year?". EarthSky. Archived from the original on February 20, 2015. Retrieved February 20, 2015. Rindler, p. 196. Christian, Eric; Samar, Safi-Harb. "How large is the Milky Way?". Archived from the original on February 2, 1999. Retrieved November 28, 2007. Hall, Shannon (May 4, 2015). "Size of the Milky Way Upgraded, Solving Galaxy Puzzle". Space.com. Archived from the original on June 7, 2015. Retrieved June 9, 2015. Ribas, I.; Jordi, C.; Vilardell, F.; Fitzpatrick, E. L.; Hilditch, R. W.; Guinan, F. Edward (2005). "First Determination of the Distance and Fundamental Properties of an Eclipsing Binary in the Andromeda Galaxy". Astrophysical Journal. 635 (1): L37–L40. arXiv:astro-ph/0511045. Bibcode:2005ApJ...635L..37R. doi:10.1086/499161. S2CID 119522151. McConnachie, A.W.; Irwin, M.J.; Ferguson, A.M.N.; Ibata, R.A.; Lewis, G.F.; Tanvir, N. (2005). "Distances and metallicities for 17 Local Group galaxies". Monthly Notices of the Royal Astronomical Society. 356 (4): 979–997. arXiv:astro-ph/0410489. Bibcode:2005MNRAS.356..979M. doi:10.1111/j.1365-2966.2004.08514.x. Janek, Vanessa (February 20, 2015). "How can space travel faster than the speed of light?". Universe Today. Archived from the original on December 16, 2021. Retrieved June 6, 2015. "Is faster-than-light travel or communication possible? Section: Expansion of the Universe". Philip Gibbs. 1997. Archived from the original on March 10, 2010. Retrieved June 6, 2015. Vardanyan, M.; Trotta, R.; Silk, J. (January 28, 2011). "Applications of Bayesian model averaging to the curvature and size of the Universe". Monthly Notices of the Royal Astronomical Society: Letters. 413 (1): L91–L95. arXiv:1101.5476. Bibcode:2011MNRAS.413L..91V. doi:10.1111/j.1745-3933.2011.01040.x. S2CID 2616287. Schreiber, Urs (June 6, 2008). "Urban Myths in Contemporary Cosmology". The n-Category Café. University of Texas at Austin. Archived from the original on July 1, 2020. Retrieved June 1, 2020. Don N. Page (2007). "Susskind's Challenge to the Hartle-Hawking No-Boundary Proposal and Possible Resolutions". Journal of Cosmology and Astroparticle Physics. 2007 (1): 004. arXiv:hep-th/0610199. Bibcode:2007JCAP...01..004P. doi:10.1088/1475-7516/2007/01/004. S2CID 17403084. Berardelli, Phil (March 25, 2010). "Galaxy Collisions Give Birth to Quasars". Science News. Archived from the original on March 25, 2022. Retrieved July 30, 2022. Riess, Adam G.; Filippenko; Challis; Clocchiatti; Diercks; Garnavich; Gilliland; Hogan; Jha; Kirshner; Leibundgut; Phillips; Reiss; Schmidt; Schommer; Smith; Spyromilio; Stubbs; Suntzeff; Tonry (1998). "Observational evidence from supernovae for an accelerating universe and a cosmological constant". Astronomical Journal. 116 (3): 1009–1038. arXiv:astro-ph/9805201. Bibcode:1998AJ....116.1009R. doi:10.1086/300499. S2CID 15640044. Perlmutter, S.; Aldering; Goldhaber; Knop; Nugent; Castro; Deustua; Fabbro; Goobar; Groom; Hook; Kim; Kim; Lee; Nunes; Pain; Pennypacker; Quimby; Lidman; Ellis; Irwin; McMahon; Ruiz-Lapuente; Walton; Schaefer; Boyle; Filippenko; Matheson; Fruchter; et al. (1999). "Measurements of Omega and Lambda from 42 high redshift supernovae". Astrophysical Journal. 517 (2): 565–586. arXiv:astro-ph/9812133. Bibcode:1999ApJ...517..565P. doi:10.1086/307221. S2CID 118910636. Serway, Raymond A.; Moses, Clement J.; Moyer, Curt A. (2004). Modern Physics. Cengage Learning. p. 21. ISBN 978-1-111-79437-8. Fraknoi, Andrew; et al. (2022). Astronomy 2e. OpenStax. p. 1017. ISBN 978-1-951-69350-3. Archived from the original on February 14, 2023. Retrieved February 14, 2023. "The Nobel Prize in Physics 2011". Archived from the original on April 17, 2015. Retrieved April 16, 2015. Overbye, Dennis (October 11, 2003). "A 'Cosmic Jerk' That Reversed the Universe". New York Times. Archived from the original on July 1, 2017. Retrieved February 20, 2017. Schutz, Bernard (2009). A First Course in General Relativity (2nd ed.). Cambridge University Press. pp. 142, 171. ISBN 978-0-521-88705-2. Mermin, N. David (2021) [2005]. It's About Time: Understanding Einstein's Relativity (Princeton Science Library paperback ed.). Princeton University Press. ISBN 978-0-691-12201-4. OCLC 1193067111. Brill, Dieter; Jacobsen, Ted (2006). "Spacetime and Euclidean geometry". General Relativity and Gravitation. 38 (4): 643–651. arXiv:gr-qc/0407022. Bibcode:2006GReGr..38..643B. CiteSeerX 10.1.1.338.7953. doi:10.1007/s10714-006-0254-9. S2CID 119067072. Wheeler, John Archibald (2010). Geons, Black Holes, and Quantum Foam: A Life in Physics. W. W. Norton & Company. ISBN 978-0-393-07948-7. Archived from the original on February 17, 2023. Retrieved February 17, 2023. Kersting, Magdalena (May 2019). "Free fall in curved spacetime – how to visualise gravity in general relativity". Physics Education. 54 (3): 035008. Bibcode:2019PhyEd..54c5008K. doi:10.1088/1361-6552/ab08f5. hdl:10852/74677. ISSN 0031-9120. S2CID 127471222. Goldstein, Herbert; Poole, Charles P.; Safko, John L. (2002). Classical Mechanics (3rd ed.). San Francisco: Addison Wesley. ISBN 0-201-31611-0. OCLC 47056311. Goodstein, Judith R. (2018). Einstein's Italian Mathematicians: Ricci, Levi-Civita, and the Birth of General Relativity. Providence, Rhode Island: American Mathematical Society. p. 143. ISBN 978-1-4704-2846-4. OCLC 1020305599. Choquet-Bruhat, Yvonne (2009). General Relativity and the Einstein Equations. Oxford: Oxford University Press. ISBN 978-0-19-155226-7. OCLC 317496332. Prescod-Weinstein, Chanda (2021). The Disordered Cosmos: A Journey into Dark Matter, Spacetime, and Dreams Deferred. New York, New York: Bold Type Books. ISBN 978-1-5417-2470-9. OCLC 1164503847. Archived from the original on February 21, 2022. Retrieved February 17, 2023. "WMAP Mission – Age of the Universe". map.gsfc.nasa.gov. Archived from the original on December 4, 2022. Retrieved February 14, 2023. Luminet, Jean-Pierre; Weeks, Jeffrey R.; Riazuelo, Alain; Lehoucq, Roland; Uzan, Jean-Philippe (October 9, 2003). "Dodecahedral space topology as an explanation for weak wide-angle temperature correlations in the cosmic microwave background". Nature (Submitted manuscript). 425 (6958): 593–595. arXiv:astro-ph/0310253. Bibcode:2003Natur.425..593L. doi:10.1038/nature01944. PMID 14534579. S2CID 4380713. Archived from the original on May 17, 2021. Retrieved August 21, 2018. Luminet, Jean-Pierre; Roukema, Boudewijn F. (1999). "Topology of the Universe: Theory and Observations". Proceedings of Cosmology School held at Cargese, Corsica, August 1998. arXiv:astro-ph/9901364. Bibcode:1999ASIC..541..117L. Harrison, Edward Robert (2000). Cosmology: the science of the universe. Cambridge University Press. pp. 447–. ISBN 978-0-521-66148-5. Archived from the original on August 26, 2016. Retrieved May 1, 2011. Liddle, Andrew R.; Lyth, David Hilary (2000). Cosmological inflation and large-scale structure. Cambridge University Press. pp. 24–. ISBN 978-0-521-57598-0. Archived from the original on December 31, 2013. Retrieved May 1, 2011. "What is the Ultimate Fate of the Universe?". National Aeronautics and Space Administration. Archived from the original on December 22, 2021. Retrieved August 23, 2015. "WMAP – Shape of the Universe". map.gsfc.nasa.gov. Archived from the original on March 31, 2019. Retrieved February 14, 2023. Roukema, Boudewijn; Buliński, Zbigniew; Szaniewska, Agnieszka; Gaudin, Nicolas E. (2008). "A test of the Poincare dodecahedral space topology hypothesis with the WMAP CMB data". Astronomy and Astrophysics. 482 (3): 747–753. arXiv:0801.0006. Bibcode:2008A&A...482..747L. doi:10.1051/0004-6361:20078777. S2CID 1616362. Aurich, Ralf; Lustig, S.; Steiner, F.; Then, H. (2004). "Hyperbolic Universes with a Horned Topology and the CMB Anisotropy". Classical and Quantum Gravity. 21 (21): 4901–4926. arXiv:astro-ph/0403597. Bibcode:2004CQGra..21.4901A. doi:10.1088/0264-9381/21/21/010. S2CID 17619026. Planck Collaboration (2014). "Planck 2013 results. XVI. Cosmological parameters". Astronomy & Astrophysics. 571: A16. arXiv:1303.5076. Bibcode:2014A&A...571A..16P. doi:10.1051/0004-6361/201321591. S2CID 118349591. "Planck reveals 'almost perfect' universe". Michael Banks. Physics World. March 21, 2013. Archived from the original on March 24, 2013. Retrieved March 21, 2013. Friederich, Simon (November 12, 2021). "Fine-Tuning". The Stanford Encyclopedia of Philosophy. Center for the Study of Language and Information (CSLI), Stanford University. Archived from the original on October 10, 2023. Retrieved February 15, 2022. Isaak, Mark, ed. (2005). "CI301: The Anthropic Principle". Index to Creationist Claims. TalkOrigins Archive. Archived from the original on July 1, 2014. Retrieved October 31, 2007. Fritzsche, Hellmut. "electromagnetic radiation | physics". Encyclopædia Britannica. p. 1. Archived from the original on August 31, 2015. Retrieved July 26, 2015. "Physics 7:Relativity, SpaceTime and Cosmology" (PDF). Physics 7:Relativity, SpaceTime and Cosmology. University of California Riverside. Archived from the original (PDF) on September 5, 2015. Retrieved July 26, 2015. "Physics – for the 21st Century". learner.org. Harvard-Smithsonian Center for Astrophysics Annenberg Learner. Archived from the original on September 7, 2015. Retrieved July 27, 2015. "Dark matter – A history shapes by dark force". Timothy Ferris. National Geographic. 2015. Archived from the original on March 4, 2016. Retrieved December 29, 2015. Redd, SPACE.com, Nola Taylor. "It's Official: The Universe Is Dying Slowly". Scientific American. Archived from the original on August 12, 2015. Retrieved August 11, 2015. Parr, Will; et al. "RIP Universe – Your Time Is Coming… Slowly | Video". Space.com. Archived from the original on August 13, 2015. Retrieved August 20, 2015. Sean Carroll, Ph.D., Caltech, 2007, The Teaching Company, Dark Matter, Dark Energy: The Dark Side of the Universe, Guidebook Part 2. p. 46, Accessed October 7, 2013, "...dark matter: An invisible, essentially collisionless component of matter that makes up about 25 percent of the energy density of the universe... it's a different kind of particle... something not yet observed in the laboratory..." Peebles, P. J. E. & Ratra, Bharat (2003). "The cosmological constant and dark energy". Reviews of Modern Physics. 75 (2): 559–606. arXiv:astro-ph/0207347. Bibcode:2003RvMP...75..559P. doi:10.1103/RevModPhys.75.559. S2CID 118961123. Mandolesi, N.; Calzolari, P.; Cortiglioni, S.; Delpino, F.; Sironi, G.; Inzani, P.; Deamici, G.; Solheim, J.-E.; Berger, L.; Partridge, R.B.; Martenis, P.L.; Sangree, C.H.; Harvey, R.C. (1986). "Large-scale homogeneity of the universe measured by the microwave background". Nature. 319 (6056): 751–753. Bibcode:1986Natur.319..751M. doi:10.1038/319751a0. S2CID 4349689. Gunn, Alistair (November 29, 2023). "How many galaxies are there in the universe? – Do astronomers know how many galaxies exist? How many can we see in the observable Universe?". BBC Sky at Night. Archived from the original on December 3, 2023. Retrieved December 2, 2023. "New Horizons spacecraft answers the question: How dark is space?". phys.org. Archived from the original on January 15, 2021. Retrieved January 15, 2021. Howell, Elizabeth (March 20, 2018). "How Many Galaxies Are There?". Space.com. Archived from the original on February 28, 2021. Retrieved March 5, 2021. Staff (2019). "How Many Stars Are There In The Universe?". European Space Agency. Archived from the original on September 23, 2019. Retrieved September 21, 2019. Marov, Mikhail Ya. (2015). "The Structure of the Universe". The Fundamentals of Modern Astrophysics. pp. 279–294. doi:10.1007/978-1-4614-8730-2_10. ISBN 978-1-4614-8729-6. Mackie, Glen (February 1, 2002). "To see the Universe in a Grain of Taranaki Sand". Centre for Astrophysics and Supercomputing. Archived from the original on August 11, 2011. Retrieved January 28, 2017. Mack, Eric (March 19, 2015). "There may be more Earth-like planets than grains of sand on all our beaches – New research contends that the Milky Way alone is flush with billions of potentially habitable planets – and that's just one sliver of the universe". CNET. Archived from the original on December 1, 2023. Retrieved December 1, 2023. T. Bovaird, T.; Lineweaver, C.H.; Jacobsen, S.K. (March 13, 2015). "Using the inclinations of Kepler systems to prioritize new Titius–Bode-based exoplanet predictions". Monthly Notices of the Royal Astronomical Society. 448 (4): 3608–3627. arXiv:1412.6230. doi:10.1093/mnras/stv221. Archived from the original on December 1, 2023. Retrieved December 1, 2023. Baker, Harry (July 11, 2021). "How many atoms are in the observable universe?". Live Science. Archived from the original on December 1, 2023. Retrieved December 1, 2023. Totani, Tomonori (February 3, 2020). "Emergence of life in an inflationary universe". Scientific Reports. 10 (1671): 1671. arXiv:1911.08092. Bibcode:2020NatSR..10.1671T. doi:10.1038/s41598-020-58060-0. PMC 6997386. PMID 32015390. "Unveiling the Secret of a Virgo Dwarf Galaxy". European Southern Observatory Press Release. ESO: 12. May 3, 2000. Bibcode:2000eso..pres...12. Archived from the original on July 13, 2015. Retrieved January 3, 2007. "Hubble's Largest Galaxy Portrait Offers a New High-Definition View". NASA. February 28, 2006. Archived from the original on May 27, 2020. Retrieved January 3, 2007. Gibney, Elizabeth (September 3, 2014). "Earth's new address: 'Solar System, Milky Way, Laniakea'". Nature. doi:10.1038/nature.2014.15819. S2CID 124323774. Archived from the original on January 7, 2019. Retrieved August 21, 2015. "Local Group". Fraser Cain. Universe Today. May 4, 2009. Archived from the original on June 21, 2018. Retrieved August 21, 2015. Devlin, Hannah; Correspondent, Science (April 20, 2015). "Astronomers discover largest known structure in the universe is ... a big hole". The Guardian. Archived from the original on February 7, 2017. Retrieved December 18, 2016. "Content of the Universe – WMAP 9yr Pie Chart". wmap.gsfc.nasa.gov. Archived from the original on September 5, 2015. Retrieved July 26, 2015. Rindler, p. 202. Liddle, Andrew (2003). An Introduction to Modern Cosmology (2nd ed.). John Wiley & Sons. ISBN 978-0-470-84835-7.. p. 2. Livio, Mario (2001). The Accelerating Universe: Infinite Expansion, the Cosmological Constant, and the Beauty of the Cosmos. John Wiley and Sons. p. 53. ISBN 978-0-471-43714-7. Archived from the original on May 13, 2021. Retrieved March 31, 2012. Peebles, P.J.E. & Ratra, Bharat (2003). "The cosmological constant and dark energy". Reviews of Modern Physics. 75 (2): 559–606. arXiv:astro-ph/0207347. Bibcode:2003RvMP...75..559P. doi:10.1103/RevModPhys.75.559. S2CID 118961123. Steinhardt, Paul J.; Turok, Neil (2006). "Why the cosmological constant is small and positive". Science. 312 (5777): 1180–1183. arXiv:astro-ph/0605173. Bibcode:2006Sci...312.1180S. doi:10.1126/science.1126231. PMID 16675662. S2CID 14178620. "Dark Energy". Hyperphysics. Archived from the original on May 27, 2013. Retrieved January 4, 2014. Carroll, Sean (2001). "The cosmological constant". Living Reviews in Relativity. 4 (1): 1. arXiv:astro-ph/0004075. Bibcode:2001LRR.....4....1C. doi:10.12942/lrr-2001-1. PMC 5256042. PMID 28179856. Archived from the original on October 13, 2006. Retrieved September 28, 2006. "Planck captures portrait of the young universe, revealing earliest light". University of Cambridge. March 21, 2013. Archived from the original on April 17, 2019. Retrieved March 21, 2013. Davies, P. (1992). The New Physics: A Synthesis. Cambridge University Press. p. 1. ISBN 978-0-521-43831-5. Archived from the original on February 3, 2021. Retrieved May 17, 2020. Persic, Massimo; Salucci, Paolo (September 1, 1992). "The baryon content of the universe". Monthly Notices of the Royal Astronomical Society. 258 (1): 14P–18P. arXiv:astro-ph/0502178. Bibcode:1992MNRAS.258P..14P. doi:10.1093/mnras/258.1.14P. ISSN 0035-8711. S2CID 17945298. Shull, J. Michael; Smith, Britton D.; Danforth, Charles W. (November 1, 2012). "The Baryon Census in a Multiphase Intergalactic Medium: 30% of the Baryons May Still Be Missing". The Astrophysical Journal. 759 (1): 23. arXiv:1112.2706. Bibcode:2012ApJ...759...23S. doi:10.1088/0004-637X/759/1/23. ISSN 0004-637X. S2CID 119295243. Archived from the original on September 21, 2023. Retrieved February 27, 2023. "Galaxy surveys have found ~10% of these baryons in collapsed objects such as galaxies, groups, and clusters [...] Of the remaining 80%–90% of cosmological baryons, approximately half can be accounted for in the low-z [intergalactic medium]" Macquart, J.-P.; Prochaska, J. X.; McQuinn, M.; Bannister, K. W.; Bhandari, S.; Day, C. K.; Deller, A. T.; Ekers, R. D.; James, C. W.; Marnoch, L.; Osłowski, S.; Phillips, C.; Ryder, S. D.; Scott, D. R.; Shannon, R. M. (May 28, 2020). "A census of baryons in the Universe from localized fast radio bursts". Nature. 581 (7809): 391–395. arXiv:2005.13161. Bibcode:2020Natur.581..391M. doi:10.1038/s41586-020-2300-2. ISSN 0028-0836. PMID 32461651. S2CID 256821489. Archived from the original on November 5, 2023. Retrieved February 27, 2023. Flowers, Paul; et al. (2019). Chemistry 2e. OpenStax. p. 14. ISBN 978-1-947-17262-3. Archived from the original on February 17, 2023. Retrieved February 17, 2023. "The Nobel Prize in Physics 2001". NobelPrize.org. Archived from the original on February 17, 2023. Retrieved February 17, 2023. Cohen-Tannoudji, Claude; Guery-Odelin, David (2011). Advances In Atomic Physics: An Overview. World Scientific. p. 684. ISBN 978-981-4390-58-3. Archived from the original on June 4, 2023. Retrieved February 17, 2023. 't Hooft, G. (1997). In search of the ultimate building blocks. Cambridge University Press. p. 6. ISBN 978-0-521-57883-7. Clayton, Donald D. (1983). Principles of Stellar Evolution and Nucleosynthesis. The University of Chicago Press. pp. 362–435. ISBN 978-0-226-10953-4. Veltman, Martinus (2003). Facts and Mysteries in Elementary Particle Physics. World Scientific. ISBN 978-981-238-149-1. Braibant, Sylvie; Giacomelli, Giorgio; Spurio, Maurizio (2012). Particles and Fundamental Interactions: An Introduction to Particle Physics (2nd ed.). Springer. pp. 1–3. ISBN 978-94-007-2463-1. Archived from the original on August 26, 2016. Retrieved January 27, 2016. Close, Frank (2012). Particle Physics: A Very Short Introduction. Oxford University Press. ISBN 978-0-19-280434-1. Mann, Adam (August 20, 2022). "What Are Elementary Particles?". Live Science. Archived from the original on August 17, 2023. Retrieved August 17, 2023. Zwiebach, Barton (2022). Mastering Quantum Mechanics: Essentials, Theory, and Applications. MIT Press. p. 31. ISBN 978-0-262-04613-8. Oerter, R. (2006). The Theory of Almost Everything: The Standard Model, the Unsung Triumph of Modern Physics (Kindle). Penguin Group. p. 2. ISBN 978-0-13-236678-6. Onyisi, P. (October 23, 2012). "Higgs boson FAQ". University of Texas ATLAS group. Archived from the original on October 12, 2013. Retrieved January 8, 2013. Strassler, M. (October 12, 2012). "The Higgs FAQ 2.0". ProfMattStrassler.com. Archived from the original on October 12, 2013. Retrieved January 8, 2013. "[Q] Why do particle physicists care so much about the Higgs particle? [A] Well, actually, they don't. What they really care about is the Higgs field, because it is so important. [emphasis in original]" Weinberg, Steven (2011). Dreams of a Final Theory: The Scientist's Search for the Ultimate Laws of Nature. Knopf Doubleday Publishing Group. ISBN 978-0-307-78786-6. Allday, Jonathan (2002). Quarks, Leptons and the Big Bang (2nd ed.). IOP Publishing. ISBN 978-0-7503-0806-9. "Lepton (physics)". Encyclopædia Britannica. Archived from the original on May 11, 2015. Retrieved September 29, 2010. Harari, H. (1977). "Beyond charm". In Balian, R.; Llewellyn-Smith, C.H. (eds.). Weak and Electromagnetic Interactions at High Energy, Les Houches, France, Jul 5 – Aug 14, 1976. Les Houches Summer School Proceedings. Vol. 29. North-Holland. p. 613. Harari H. (1977). "Three generations of quarks and leptons" (PDF). In E. van Goeler; Weinstein R. (eds.). Proceedings of the XII Rencontre de Moriond. p. 170. SLAC-PUB-1974. Archived (PDF) from the original on May 13, 2020. Retrieved May 29, 2020. "Experiment confirms famous physics model" (Press release). MIT News Office. April 18, 2007. Archived from the original on July 5, 2013. Retrieved June 2, 2015. "Thermal history of the universe and early growth of density fluctuations" (PDF). Guinevere Kauffmann. Max Planck Institute for Astrophysics. Archived (PDF) from the original on August 21, 2016. Retrieved January 6, 2016. "First few minutes". Eric Chaisson. Harvard Smithsonian Center for Astrophysics. Archived from the original on December 4, 2013. Retrieved January 6, 2016. "Timeline of the Big Bang". The physics of the Universe. Archived from the original on March 30, 2020. Retrieved January 6, 2016. Zeilik, Michael; Gregory, Stephen A. (1998). "25-2". Introductory Astronomy & Astrophysics (4th ed.). Saunders College Publishing. ISBN 978-0-03-006228-5. Raine & Thomas (2001, p. 12) Raine & Thomas (2001, p. 66) Friedmann, A. (1922). "Über die Krümmung des Raumes" (PDF). Zeitschrift für Physik. 10 (1): 377–386. Bibcode:1922ZPhy...10..377F. doi:10.1007/BF01332580. S2CID 125190902. Archived (PDF) from the original on May 15, 2016. Retrieved August 13, 2015. Raine & Thomas (2001, pp. 122–123) Raine & Thomas (2001, p. 70) Raine & Thomas (2001, p. 84) Raine & Thomas (2001, pp. 88, 110–113) Munitz, M. K. (1959). "One Universe or Many?". Journal of the History of Ideas. 12 (2): 231–255. doi:10.2307/2707516. JSTOR 2707516. Linde, A. (1986). "Eternal chaotic inflation". Mod. Phys. Lett. A. 1 (2): 81–85. Bibcode:1986MPLA....1...81L. doi:10.1142/S0217732386000129. S2CID 123472763. Archived from the original on April 17, 2019. Retrieved August 6, 2017. Linde, A. (1986). "Eternally existing self-reproducing chaotic inflationary Universe" (PDF). Phys. Lett. B. 175 (4): 395–400. Bibcode:1986PhLB..175..395L. doi:10.1016/0370-2693(86)90611-8. Archived (PDF) from the original on November 27, 2013. Retrieved March 17, 2011. Everett, Hugh (1957). "Relative State Formulation of Quantum Mechanics". Reviews of Modern Physics. 29 (3): 454–462. Bibcode:1957RvMP...29..454E. doi:10.1103/RevModPhys.29.454. S2CID 17178479. Ball, Philip (February 17, 2015). "Too many worlds". Aeon.co. Archived from the original on September 27, 2021. Retrieved September 23, 2021. Peres, Asher (1995). Quantum Theory: Concepts and Methods. Kluwer Academic Publishers. p. 374. ISBN 0-7923-2549-4. Kent, Adrian (February 2015). "Does it Make Sense to Speak of Self-Locating Uncertainty in the Universal Wave Function? Remarks on Sebens and Carroll". Foundations of Physics. 45 (2): 211–217. arXiv:1408.1944. Bibcode:2015FoPh...45..211K. doi:10.1007/s10701-014-9862-5. ISSN 0015-9018. S2CID 118471198. Schlosshauer, Maximilian; Kofler, Johannes; Zeilinger, Anton (August 1, 2013). "A snapshot of foundational attitudes toward quantum mechanics". Studies in History and Philosophy of Science Part B: Studies in History and Philosophy of Modern Physics. 44 (3): 222–230. arXiv:1301.1069. Bibcode:2013SHPMP..44..222S. doi:10.1016/j.shpsb.2013.04.004. ISSN 1355-2198. S2CID 55537196. Mermin, N. David (July 1, 2012). "Commentary: Quantum mechanics: Fixing the shifty split". Physics Today. 65 (7): 8–10. Bibcode:2012PhT....65g...8M. doi:10.1063/PT.3.1618. ISSN 0031-9228. "New interpretations appear every year. None ever disappear." Cabello, Adán (2017). "Interpretations of quantum theory: A map of madness". In Lombardi, Olimpia; Fortin, Sebastian; Holik, Federico; López, Cristian (eds.). What is Quantum Information?. Cambridge University Press. pp. 138–143. arXiv:1509.04711. Bibcode:2015arXiv150904711C. doi:10.1017/9781316494233.009. ISBN 9781107142114. S2CID 118419619. Garriga, Jaume; Vilenkin, Alexander (2007). "Many Worlds in One". Physical Review D. 64 (4): 043511. arXiv:gr-qc/0102010v2. doi:10.1103/PhysRevD.64.043511. S2CID 119000743. Tegmark, Max (2003). "Parallel universes. Not just a staple of science fiction, other universes are a direct implication of cosmological observations". Scientific American. 288 (5): 40–51. arXiv:astro-ph/0302131. Bibcode:2003SciAm.288e..40T. doi:10.1038/scientificamerican0503-40. PMID 12701329. Tegmark, Max (2003). "Parallel Universes". Scientific American. 288 (5): 40–51. arXiv:astro-ph/0302131. Bibcode:2003SciAm.288e..40T. doi:10.1038/scientificamerican0503-40. PMID 12701329. Gil, Francisco José Soler; Alfonseca, Manuel (2013). "About the Infinite Repetition of Histories in Space". Theoria: An International Journal for Theory, History and Foundations of Science. 29 (3): 361. arXiv:1301.5295. doi:10.1387/theoria.9951. hdl:10486/664735. S2CID 52996408. Ellis, G. F. (2011). "Does the Multiverse Really Exist?". Scientific American. 305 (2): 38–43. Bibcode:2011SciAm.305a..38E. doi:10.1038/scientificamerican0811-38. PMID 21827123. Moskowitz, Clara (August 12, 2011). "Weird! Our Universe May Be a 'Multiverse,' Scientists Say". livescience. Archived from the original on May 5, 2015. Retrieved May 4, 2015. Gernet, J. (1993–1994). "Space and time: Science and religion in the encounter between China and Europe". Chinese Science. Vol. 11. pp. 93–102. Blandford R. D. (2015). "A century of general relativity: Astrophysics and cosmology". Science. 347 (6226): 1103–1108. Bibcode:2015Sci...347.1103B. doi:10.1126/science.aaa4033. PMID 25745165. S2CID 30364122. Leeming, David A. (2010). Creation Myths of the World. ABC-CLIO. p. xvii. ISBN 978-1-59884-174-9. "In common usage the word 'myth' refers to narratives or beliefs that are untrue or merely fanciful; the stories that make up national or ethnic mythologies describe characters and events that common sense and experience tell us are impossible. Nevertheless, all cultures celebrate such myths and attribute to them various degrees of literal or symbolic truth." Eliade, Mircea (1964). Myth and Reality (Religious Traditions of the World). Allen & Unwin. ISBN 978-0-04-291001-7. Leonard, Scott A.; McClure, Michael (2004). Myth and Knowing: An Introduction to World Mythology. McGraw-Hill. ISBN 978-0-7674-1957-4. (Henry Gravrand, "La civilisation Sereer -Pangool") [in] Universität Frankfurt am Main, Frobenius-Institut, Deutsche Gesellschaft für Kulturmorphologie, Frobenius Gesellschaft, "Paideuma: Mitteilungen zur Kulturkunde, Volumes 43–44", F. Steiner (1997), pp. 144–145, ISBN 3-515-02842-0 Young, Louise B. (1993). The Unfinished Universe. Oxford University Press. p. 21. ISBN 978-0-195-08039-1. OCLC 26399171. Graham, Daniel W. (September 3, 2019). "Heraclitus". In Zalta, Edward N. (ed.). Stanford Encyclopedia of Philosophy. Palmer, John (October 19, 2020). "Parmenides". In Zalta, Edward N. (ed.). Stanford Encyclopedia of Philosophy. Palmer, John (April 8, 2021). "Zeno of Elea". In Zalta, Edward N. (ed.). Stanford Encyclopedia of Philosophy. Dowden, Bradley. "Zeno's Paradoxes". Internet Encyclopedia of Philosophy. Will Durant, Our Oriental Heritage: "Two systems of Hindu thought propound physical theories suggestively similar to those of Greece. Kanada, founder of the Vaisheshika philosophy, held that the world is composed of atoms as many in kind as the various elements. The Jains more nearly approximated to Democritus by teaching that all atoms were of the same kind, producing different effects by diverse modes of combinations. Kanada believed light and heat to be varieties of the same substance; Udayana taught that all heat comes from the Sun; and Vachaspati, like Newton, interpreted light as composed of minute particles emitted by substances and striking the eye." Stcherbatsky, F. Th. (1930, 1962), Buddhist Logic, Volume 1, p. 19, Dover, New York: "The Buddhists denied the existence of substantial matter altogether. Movement consists for them of moments, it is a staccato movement, momentary flashes of a stream of energy... "Everything is evanescent",... says the Buddhist, because there is no stuff... Both systems [Sānkhya, and later Indian Buddhism] share in common a tendency to push the analysis of existence up to its minutest, last elements which are imagined as absolute qualities, or things possessing only one unique quality. They are called "qualities" (guna-dharma) in both systems in the sense of absolute qualities, a kind of atomic, or intra-atomic, energies of which the empirical things are composed. Both systems, therefore, agree in denying the objective reality of the categories of Substance and Quality,... and of the relation of Inference uniting them. There is in Sānkhya philosophy no separate existence of qualities. What we call quality is but a particular manifestation of a subtle entity. To every new unit of quality corresponds a subtle quantum of matter which is called guna, "quality", but represents a subtle substantive entity. The same applies to early Buddhism where all qualities are substantive... or, more precisely, dynamic entities, although they are also called dharmas ('qualities')." Viney, Donald Wayne (1985). "The Cosmological Argument". Charles Hartshorne and the Existence of God. SUNY Press. pp. 65–68. ISBN 978-0-87395-907-0. Pearsall, Judy (1998). The New Oxford Dictionary Of English (1st ed.). Oxford: Clarendon Press. p. 1341. ISBN 978-0-19-861263-6. Edwards, Paul (1967). Encyclopedia of Philosophy. New York: Macmillan. p. 34. Encyclopedia of Philosophy ed. Paul Edwards. New York: Macmillan and Free Press. 1967. p. 34. Reid-Bowen, Paul (April 15, 2016). Goddess as Nature: Towards a Philosophical Thealogy. Taylor & Francis. p. 70. ISBN 9781317126348. Lindberg, David C. (2007). The beginnings of Western science: the European Scientific tradition in philosophical, religious, and institutional context (2nd ed.). University of Chicago Press. p. 12. ISBN 9780226482057. Grant, Edward (2007). "Ancient Egypt to Plato". A History of Natural Philosophy: From the Ancient World to the Nineteenth Century. New York: Cambridge University Press. pp. 1–26. ISBN 978-0-521-68957-1. Horowitz, Wayne (1988). "The Babylonian Map of the World". Iraq. 50: 147–165. doi:10.2307/4200289. JSTOR 4200289. S2CID 190703581. Keel, Othmar (1997). The Symbolism of the Biblical World. Eisenbrauns. pp. 20–22. ISBN 978-1-575-06014-9. Archived from the original on March 13, 2024. Retrieved February 26, 2023. Wright, Larry (August 1973). "The astronomy of Eudoxus: Geometry or physics?". Studies in History and Philosophy of Science. 4 (2): 165–172. Bibcode:1973SHPSA...4..165W. doi:10.1016/0039-3681(73)90002-2. Archived from the original on March 15, 2023. Retrieved February 27, 2023. Dicati, Renato (2013). "The Ancients' Astronomy". Stamping Through Astronomy. Milano: Springer Milan. pp. 19–55. doi:10.1007/978-88-470-2829-6_2. ISBN 978-88-470-2828-9. Archived from the original on March 13, 2024. Retrieved February 27, 2023. Aristotle; Forster, E. S.; Dobson, J. F. (1914). De Mundo. Oxford: The Clarendon Press. p. 2. Goldstein, Bernard R. (1997). "Saving the phenomena: the background to Ptolemy's planetary theory". Journal for the History of Astronomy. 28 (1): 1–12. Bibcode:1997JHA....28....1G. doi:10.1177/002182869702800101. S2CID 118875902. Boyer, C. (1968) A History of Mathematics. Wiley, p. 54. Heath, Thomas (2013). Aristarchus of Samos, the Ancient Copernicus: A History of Greek Astronomy to Aristarchus, Together with Aristarchus's Treatise on the Sizes and Distances of the Sun and Moon. Cambridge University Press. p. 302. ISBN 978-1-108-06233-6. Archived from the original on March 13, 2024. Retrieved February 26, 2023. Kolkata, James J. (2015). Elementary Cosmology: From Aristotle's Universe to the Big Bang and Beyond. IOP Publishing. doi:10.1088/978-1-6817-4100-0ch4. ISBN 978-1-68174-100-0. Archived from the original on June 5, 2018. Retrieved February 27, 2023. Neugebauer, Otto E. (1945). "The History of Ancient Astronomy Problems and Methods". Journal of Near Eastern Studies. 4 (1): 166–173. doi:10.1086/370729. JSTOR 595168. S2CID 162347339. "the Chaldaean Seleucus from Seleucia" Sarton, George (1955). "Chaldaean Astronomy of the Last Three Centuries B. C.". Journal of the American Oriental Society. 75 (3): 166–173 [169]. doi:10.2307/595168. JSTOR 595168. "the heliocentrical astronomy invented by Aristarchos of Samos and still defended a century later by Seleucos the Babylonian" William P. D. Wightman (1951, 1953), The Growth of Scientific Ideas, Yale University Press. p. 38, where Wightman calls him Seleukos the Chaldean. Lucio Russo, Flussi e riflussi, Feltrinelli, Milano, Italy, 2003, ISBN 88-07-10349-4. Bartel (1987, p. 527) Bartel (1987, pp. 527–529) Bartel (1987, pp. 534–537) Nasr, Seyyed H. (1993) [1964]. An Introduction to Islamic Cosmological Doctrines (2nd ed.). 1st edition by Harvard University Press, 2nd edition by State University of New York Press. pp. 135–136. ISBN 978-0-7914-1515-3. Frautschi, Steven C.; Olenick, Richard P.; Apostol, Tom M.; Goodstein, David L. (2007). The Mechanical Universe: Mechanics and Heat (Advanced ed.). Cambridge [Cambridgeshire]: Cambridge University Press. p. 58. ISBN 978-0-521-71590-4. OCLC 227002144. Misner, Thorne and Wheeler, p. 754. Ālī, Ema Ākabara. Science in the Quran. Vol. 1. Malik Library. p. 218. Ragep, F. Jamil (2001). "Tusi and Copernicus: The Earth's Motion in Context". Science in Context. 14 (1–2): 145–163. doi:10.1017/s0269889701000060. S2CID 145372613. Misner, Thorne and Wheeler, pp. 755–756. Misner, Thorne and Wheeler, p. 756. de Cheseaux JPL (1744). Traité de la Comète. Lausanne. pp. 223ff.. Reprinted as Appendix II in Dickson, F. P. (1969). The Bowl of Night: The Physical Universe and Scientific Thought. Cambridge, Massachusetts: M.I.T. Press. ISBN 978-0-262-54003-2. Olbers HWM (1826). "Unknown title". Bode's Jahrbuch. 111.. Reprinted as Appendix I in Dickson, F. P. (1969). The Bowl of Night: The Physical Universe and Scientific Thought. Cambridge, Massachusetts: M.I.T. Press. ISBN 978-0-262-54003-2. Jeans, J. H. (1902). "The Stability of a Spherical Nebula". Philosophical Transactions of the Royal Society A. 199 (312–320): 1–53. Bibcode:1902RSPTA.199....1J. doi:10.1098/rsta.1902.0012. JSTOR 90845. Misner, Thorne and Wheeler, p. 757. Jones, Kenneth Glyn (February 1971). "The Observational Basis for Kant's Cosmogony: A Critical Analysis". Journal for the History of Astronomy. 2 (1): 29–34. Bibcode:1971JHA.....2...29J. doi:10.1177/002182867100200104. ISSN 0021-8286. S2CID 126269712. Archived from the original on February 27, 2023. Retrieved February 27, 2023. Smith, Robert W. (February 2008). "Beyond the Galaxy: The Development of Extragalactic Astronomy 1885–1965, Part 1". Journal for the History of Astronomy. 39 (1): 91–119. Bibcode:2008JHA....39...91S. doi:10.1177/002182860803900106. ISSN 0021-8286. S2CID 117430789. Archived from the original on February 27, 2023. Retrieved February 27, 2023. Sharov, Aleksandr Sergeevich; Novikov, Igor Dmitrievich (1993). Edwin Hubble, the discoverer of the big bang universe. Cambridge University Press. p. 34. ISBN 978-0-521-41617-7. Archived from the original on June 23, 2013. Retrieved December 31, 2011. Einstein, Albert (1917). "Kosmologische Betrachtungen zur allgemeinen Relativitätstheorie". Preussische Akademie der Wissenschaften, Sitzungsberichte. 1917. (part 1): 142–152. Bibliography Bartel, Leendert van der Waerden (1987). "The Heliocentric System in Greek, Persian and Hindu Astronomy". Annals of the New York Academy of Sciences. 500 (1): 525–545. Bibcode:1987NYASA.500..525V. doi:10.1111/j.1749-6632.1987.tb37224.x. S2CID 222087224. Landau L, Lifshitz E (1975). The Classical Theory of Fields (Course of Theoretical Physics). Vol. 2 (4th ed.). New York: Pergamon Press. pp. 358–397. ISBN 978-0-08-018176-9. Liddell, H. G. & Scott, R. (1968). A Greek-English Lexicon. Oxford University Press. ISBN 978-0-19-864214-5. Misner; C.W.; Thorne; Kip; Wheeler; J.A. (1973). Gravitation. San Francisco: W. H. Freeman. pp. 703–816. ISBN 978-0-7167-0344-0. Raine, D. J.; Thomas, E. G. (2001). An Introduction to the Science of Cosmology. Institute of Physics Publishing. Rindler, W. (1977). Essential Relativity: Special, General, and Cosmological. New York: Springer Verlag. pp. 193–244. ISBN 978-0-387-10090-6. Rees, Martin, ed. (2012). Smithsonian Universe (2nd ed.). London: Dorling Kindersley. ISBN 978-0-7566-9841-6. External links Listen to this article (4 parts, 1 hour and 13 minutes) Duration: 13 minutes and 33 seconds.13:33 Duration: 21 minutes and 49 seconds.21:49 Duration: 25 minutes and 42 seconds.25:42 Duration: 11 minutes and 33 seconds.11:33 Spoken Wikipedia icon These audio files were created from a revision of this article dated 13 June 2012, and do not reflect subsequent edits. (Audio help · More spoken articles) NASA/IPAC Extragalactic Database (NED) / (NED-Distances). There are about 1082 atoms in the observable universe – LiveScience, July 2021. This is why we will never know everything about our universe – Forbes, May 2019. Universe at Wikipedia's sister projects: Definitions from Wiktionary Media from Commons Quotations from Wikiquote vte Cosmology Background Age of the universe Big Bang Chronology of the universe Universe Observable universe History of cosmological theories Discovery of cosmic microwave background History of the Big Bang theory Religious interpretations of the Big Bang Timeline of cosmological theories Past universe Cosmic microwave background Cosmic neutrino background Gravitational wave background Inflation Nucleosynthesis Present universe FLRW metric Friedmann equations Hubble's law Expansion of the universe Accelerating expansion Redshift Future universe Future of an expanding universe Ultimate fate of the universe Components Dark energy Dark fluid Dark matter Quintessence Lambda-CDM model Structure formation Galaxy filament Galaxy formation Large quasar group Large-scale structure Reionization Shape of the universe Structure formation Experiments 2dF 6dF BOOMERanG COBE Illustris project Observational cosmology Planck SDSS WMAP astronomy portal vte Location of Earth Included Earth → Solar System → Local Interstellar Cloud → Local Bubble → Gould Belt → Orion Arm → Milky Way → Milky Way subgroup → Local Group → Local Sheet → Virgo Supercluster → Laniakea Supercluster → Local Hole → Observable universe → Universe Each arrow (→) may be read as "within" or "part of". Related Cosmic View (1957 book) To the Moon and Beyond (1964 film) Cosmic Zoom (1968 film) Powers of Ten (1968 and 1977 films) Cosmic Voyage (1996 documentary) Cosmic Eye (2012) History of the center of the Universe Order of magnitude Pisces–Cetus Supercluster Complex Astronomy portal Space portal vte The fundamental interactions of physics Physical forces Strong interaction fundamental residual Electroweak interaction weak interaction electromagnetism Gravitation Radiations Electromagnetic radiation Gravitational radiation Hypothetical forces Fifth force Quintessence Weak gravity conjecture Glossary of physics Particle physics Philosophy of physics Universe Weakless universe vte Elements of nature Universe Space Time Energy Matter particles chemical elements Change Earth Earth science History (geological) Structure Geology Plate tectonics Oceans Gaia hypothesis Future Weather Meteorology Atmosphere (Earth) Climate Clouds Moonlight Rain Snow Sunlight Tides Wind tornado tropical cyclone Natural environment Ecology Ecosystem Field Radiation Wilderness Wildfires Life Origin (abiogenesis) Evolutionary history Biosphere Hierarchy Biology (astrobiology) Biodiversity Organism Eukaryota flora plants fauna animals fungi protista Prokaryotes archaea bacteria Viruses Category Portals: icon Stars Spaceflight Solar System icon Science Authority control databases Edit this at Wikidata International FAST National France BnF data Germany Israel United States Japan Czech Republic Korea Other Encyclopedia of Modern Ukraine İslâm Ansiklopedisi Categories: UniverseAstronomical dynamical systemsConcepts in astronomyEnvironmentsMain topic articlesPhysical cosmologyScience Article Talk Read View source View history Tools Page semi-protected From Wikipedia, the free encyclopedia For other uses, see Science (disambiguation). For a topical guide, see Outline of science. Part of a series on Science icon Science portalOutlineCategoryIndexGlossaryDisambiguationHistoryLiteraturePhilosophy Fields (Outline / List) Intrascientific fields Applied sciencesFormal sciences MathematicalComputerInterdisciplinary sciencesNatural sciences PhysicalLifeEnvironmentalSocial sciences CulturalEconomicalHumanPolitical Extrascientific fields ArtsCommunication studiesCraftFuturologyHistoryHumanitiesKnowledge managementLanguage studiesLawLiberal artsLiteratureMusicPhilosophyPolemologyProfessionsReligionResearch and developmentStrategic studiesUrban studiesVocational education Scientific integrity ReproducibilityCognitive biasLogical fallacyResearch ethics Instruments Science communicationScience educationResearch fundingScientific methodScience policyScientistResearchTechnology This is a subseries on philosophy. In order to explore related topics, please visit navigation. vte Science is a rigorous, systematic endeavor that builds and organizes knowledge in the form of testable explanations and predictions about the world.[1][2] Modern science is typically divided into three major branches:[3] the natural sciences (e.g., physics, chemistry, and biology), which study the physical world; the social sciences (e.g., economics, psychology, and sociology), which study individuals and societies;[4][5] and the formal sciences (e.g., logic, mathematics, and theoretical computer science), which study formal systems, governed by axioms and rules.[6][7] There is disagreement whether the formal sciences are science disciplines,[8][9][10] as they do not rely on empirical evidence.[11][9] Applied sciences are disciplines that use scientific knowledge for practical purposes, such as in engineering and medicine.[12][13][14] The history of science spans the majority of the historical record, with the earliest written records of identifiable predecessors to modern science dating to Bronze Age Egypt and Mesopotamia from around 3000 to 1200 BCE. Their contributions to mathematics, astronomy, and medicine entered and shaped the Greek natural philosophy of classical antiquity, whereby formal attempts were made to provide explanations of events in the physical world based on natural causes, while further advancements, including the introduction of the Hindu–Arabic numeral system, were made during the Golden Age of India.[15]: 12 [16][17][18] Scientific research deteriorated in these regions after the fall of the Western Roman Empire during the Early Middle Ages (400 to 1000 CE), but in the Medieval renaissances (Carolingian Renaissance, Ottonian Renaissance and the Renaissance of the 12th century) scholarship flourished again. Some Greek manuscripts lost in Western Europe were preserved and expanded upon in the Middle East during the Islamic Golden Age,[19] along with the later efforts of Byzantine Greek scholars who brought Greek manuscripts from the dying Byzantine Empire to Western Europe at the start of the Renaissance. The recovery and assimilation of Greek works and Islamic inquiries into Western Europe from the 10th to 13th century revived "natural philosophy",[20][21][22] which was later transformed by the Scientific Revolution that began in the 16th century[23] as new ideas and discoveries departed from previous Greek conceptions and traditions.[24][25] The scientific method soon played a greater role in knowledge creation and it was not until the 19th century that many of the institutional and professional features of science began to take shape,[26][27] along with the changing of "natural philosophy" to "natural science".[28] New knowledge in science is advanced by research from scientists who are motivated by curiosity about the world and a desire to solve problems.[29][30] Contemporary scientific research is highly collaborative and is usually done by teams in academic and research institutions,[31] government agencies, and companies.[32][33] The practical impact of their work has led to the emergence of science policies that seek to influence the scientific enterprise by prioritizing the ethical and moral development of commercial products, armaments, health care, public infrastructure, and environmental protection. Etymology The word science has been used in Middle English since the 14th century in the sense of "the state of knowing". The word was borrowed from the Anglo-Norman language as the suffix -cience, which was borrowed from the Latin word scientia, meaning "knowledge, awareness, understanding". It is a noun derivative of the Latin sciens meaning "knowing", and undisputedly derived from the Latin sciō, the present participle scīre, meaning "to know".[34] There are many hypotheses for science's ultimate word origin. According to Michiel de Vaan, Dutch linguist and Indo-Europeanist, sciō may have its origin in the Proto-Italic language as *skije- or *skijo- meaning "to know", which may originate from Proto-Indo-European language as *skh1-ie, *skh1-io, meaning "to incise". The Lexikon der indogermanischen Verben proposed sciō is a back-formation of nescīre, meaning "to not know, be unfamiliar with", which may derive from Proto-Indo-European *sekH- in Latin secāre, or *skh2-, from *sḱʰeh2(i)- meaning "to cut".[35] In the past, science was a synonym for "knowledge" or "study", in keeping with its Latin origin. A person who conducted scientific research was called a "natural philosopher" or "man of science".[36] In 1834, William Whewell introduced the term scientist in a review of Mary Somerville's book On the Connexion of the Physical Sciences,[37] crediting it to "some ingenious gentleman" (possibly himself).[38] History Main article: History of science Early history Main article: History of science in early cultures Clay tablet with markings, three columns for numbers and one for ordinals The Plimpton 322 tablet by the Babylonians records Pythagorean triples, written in about 1800 BCE Science has no single origin. Rather, systematic methods emerged gradually over the course of tens of thousands of years,[39][40] taking different forms around the world, and few details are known about the very earliest developments. Women likely played a central role in prehistoric science,[41] as did religious rituals.[42] Some scholars use the term "protoscience" to label activities in the past that resemble modern science in some but not all features;[43][44][45] however, this label has also been criticized as denigrating,[46] or too suggestive of presentism, thinking about those activities only in relation to modern categories.[47] Direct evidence for scientific processes becomes clearer with the advent of writing systems in early civilizations like Ancient Egypt and Mesopotamia, creating the earliest written records in the history of science in around 3000 to 1200 BCE.[15]: 12–15 [16] Although the words and concepts of "science" and "nature" were not part of the conceptual landscape at the time, the ancient Egyptians and Mesopotamians made contributions that would later find a place in Greek and medieval science: mathematics, astronomy, and medicine.[48][15]: 12  From the 3rd millennium BCE, the ancient Egyptians developed a decimal numbering system,[49] solved practical problems using geometry,[50] and developed a calendar.[51] Their healing therapies involved drug treatments and the supernatural, such as prayers, incantations, and rituals.[15]: 9  The ancient Mesopotamians used knowledge about the properties of various natural chemicals for manufacturing pottery, faience, glass, soap, metals, lime plaster, and waterproofing.[52] They studied animal physiology, anatomy, behavior, and astrology for divinatory purposes.[53] The Mesopotamians had an intense interest in medicine and the earliest medical prescriptions appeared in Sumerian during the Third Dynasty of Ur.[52][54] They seem to have studied scientific subjects which had practical or religious applications and had little interest in satisfying curiosity.[52] Classical antiquity Main article: Science in classical antiquity Framed mosaic of philosophers gathering around and conversing Plato's Academy mosaic, made between 100 BCE to 79 AD, shows many Greek philosophers and scholars In classical antiquity, there is no real ancient analog of a modern scientist. Instead, well-educated, usually upper-class, and almost universally male individuals performed various investigations into nature whenever they could afford the time.[55] Before the invention or discovery of the concept of phusis or nature by the pre-Socratic philosophers, the same words tend to be used to describe the natural "way" in which a plant grows,[56] and the "way" in which, for example, one tribe worships a particular god. For this reason, it is claimed that these men were the first philosophers in the strict sense and the first to clearly distinguish "nature" and "convention".[57] The early Greek philosophers of the Milesian school, which was founded by Thales of Miletus and later continued by his successors Anaximander and Anaximenes, were the first to attempt to explain natural phenomena without relying on the supernatural.[58] The Pythagoreans developed a complex number philosophy[59]: 467–68  and contributed significantly to the development of mathematical science.[59]: 465  The theory of atoms was developed by the Greek philosopher Leucippus and his student Democritus.[60][61] Later, Epicurus would develop a full natural cosmology based on atomism, and would adopt a "canon" (ruler, standard) which established physical criteria or standards of scientific truth.[62] The Greek doctor Hippocrates established the tradition of systematic medical science[63][64] and is known as "The Father of Medicine".[65] A turning point in the history of early philosophical science was Socrates' example of applying philosophy to the study of human matters, including human nature, the nature of political communities, and human knowledge itself. The Socratic method as documented by Plato's dialogues is a dialectic method of hypothesis elimination: better hypotheses are found by steadily identifying and eliminating those that lead to contradictions. The Socratic method searches for general commonly-held truths that shape beliefs and scrutinizes them for consistency.[66] Socrates criticized the older type of study of physics as too purely speculative and lacking in self-criticism.[67] Aristotle in the 4th century BCE created a systematic program of teleological philosophy.[68] In the 3rd century BCE, Greek astronomer Aristarchus of Samos was the first to propose a heliocentric model of the universe, with the Sun at the center and all the planets orbiting it.[69] Aristarchus's model was widely rejected because it was believed to violate the laws of physics,[69] while Ptolemy's Almagest, which contains a geocentric description of the Solar System, was accepted through the early Renaissance instead.[70][71] The inventor and mathematician Archimedes of Syracuse made major contributions to the beginnings of calculus.[72] Pliny the Elder was a Roman writer and polymath, who wrote the seminal encyclopedia Natural History.[73][74][75] Positional notation for representing numbers likely emerged between the 3rd and 5th centuries CE along Indian trade routes. This numeral system made efficient arithmetic operations more accessible and would eventually become standard for mathematics worldwide.[76] Middle Ages Main article: History of science § Middle Ages Picture of a peacock on very old paper The first page of Vienna Dioscurides depicts a peacock, made in the 6th century Due to the collapse of the Western Roman Empire, the 5th century saw an intellectual decline and knowledge of Greek conceptions of the world deteriorated in Western Europe.[15]: 194  During the period, Latin encyclopedists such as Isidore of Seville preserved the majority of general ancient knowledge.[77] In contrast, because the Byzantine Empire resisted attacks from invaders, they were able to preserve and improve prior learning.[15]: 159  John Philoponus, a Byzantine scholar in the 500s, started to question Aristotle's teaching of physics, introducing the theory of impetus.[15]: 307, 311, 363, 402  His criticism served as an inspiration to medieval scholars and Galileo Galilei, who extensively cited his works ten centuries later.[15]: 307–308 [78] During late antiquity and the early Middle Ages, natural phenomena were mainly examined via the Aristotelian approach. The approach includes Aristotle's four causes: material, formal, moving, and final cause.[79] Many Greek classical texts were preserved by the Byzantine empire and Arabic translations were done by groups such as the Nestorians and the Monophysites. Under the Caliphate, these Arabic translations were later improved and developed by Arabic scientists.[80] By the 6th and 7th centuries, the neighboring Sassanid Empire established the medical Academy of Gondeshapur, which is considered by Greek, Syriac, and Persian physicians as the most important medical center of the ancient world.[81] The House of Wisdom was established in Abbasid-era Baghdad, Iraq,[82] where the Islamic study of Aristotelianism flourished[83] until the Mongol invasions in the 13th century. Ibn al-Haytham, better known as Alhazen, used controlled experiment in his optical study.[a][85][86] Avicenna's compilation of the Canon of Medicine, a medical encyclopedia, is considered to be one of the most important publications in medicine and was used until the 18th century.[87] By the eleventh century, most of Europe had become Christian,[15]: 204  and in 1088, the University of Bologna emerged as the first university in Europe.[88] As such, demand for Latin translation of ancient and scientific texts grew,[15]: 204  a major contributor to the Renaissance of the 12th century. Renaissance scholasticism in western Europe flourished, with experiments done by observing, describing, and classifying subjects in nature.[89] In the 13th century, medical teachers and students at Bologna began opening human bodies, leading to the first anatomy textbook based on human dissection by Mondino de Luzzi.[90] Renaissance Main articles: Scientific Revolution and Science in the Renaissance Drawing of planets' orbit around the Sun Drawing of the heliocentric model as proposed by the Copernicus's De revolutionibus orbium coelestium New developments in optics played a role in the inception of the Renaissance, both by challenging long-held metaphysical ideas on perception, as well as by contributing to the improvement and development of technology such as the camera obscura and the telescope. At the start of the Renaissance, Roger Bacon, Vitello, and John Peckham each built up a scholastic ontology upon a causal chain beginning with sensation, perception, and finally apperception of the individual and universal forms of Aristotle.[84]: Book I  A model of vision later known as perspectivism was exploited and studied by the artists of the Renaissance. This theory uses only three of Aristotle's four causes: formal, material, and final.[91] In the sixteenth century, Nicolaus Copernicus formulated a heliocentric model of the Solar System, stating that the planets revolve around the Sun, instead of the geocentric model where the planets and the Sun revolve around the Earth. This was based on a theorem that the orbital periods of the planets are longer as their orbs are farther from the center of motion, which he found not to agree with Ptolemy's model.[92] Johannes Kepler and others challenged the notion that the only function of the eye is perception, and shifted the main focus in optics from the eye to the propagation of light.[91][93] Kepler is best known, however, for improving Copernicus' heliocentric model through the discovery of Kepler's laws of planetary motion. Kepler did not reject Aristotelian metaphysics and described his work as a search for the Harmony of the Spheres.[94] Galileo had made significant contributions to astronomy, physics and engineering. However, he became persecuted after Pope Urban VIII sentenced him for writing about the heliocentric model.[95] The printing press was widely used to publish scholarly arguments, including some that disagreed widely with contemporary ideas of nature.[96] Francis Bacon and René Descartes published philosophical arguments in favor of a new type of non-Aristotelian science. Bacon emphasized the importance of experiment over contemplation, questioned the Aristotelian concepts of formal and final cause, promoted the idea that science should study the laws of nature and the improvement of all human life.[97] Descartes emphasized individual thought and argued that mathematics rather than geometry should be used to study nature.[98] Age of Enlightenment Main article: Science in the Age of Enlightenment see caption Title page of the 1687 first edition of Philosophiæ Naturalis Principia Mathematica by Isaac Newton At the start of the Age of Enlightenment, Isaac Newton formed the foundation of classical mechanics by his Philosophiæ Naturalis Principia Mathematica, greatly influencing future physicists.[99] Gottfried Wilhelm Leibniz incorporated terms from Aristotelian physics, now used in a new non-teleological way. This implied a shift in the view of objects: objects were now considered as having no innate goals. Leibniz assumed that different types of things all work according to the same general laws of nature, with no special formal or final causes.[100] During this time, the declared purpose and value of science became producing wealth and inventions that would improve human lives, in the materialistic sense of having more food, clothing, and other things. In Bacon's words, "the real and legitimate goal of sciences is the endowment of human life with new inventions and riches", and he discouraged scientists from pursuing intangible philosophical or spiritual ideas, which he believed contributed little to human happiness beyond "the fume of subtle, sublime or pleasing [speculation]".[101] Science during the Enlightenment was dominated by scientific societies and academies,[102] which had largely replaced universities as centers of scientific research and development. Societies and academies were the backbones of the maturation of the scientific profession. Another important development was the popularization of science among an increasingly literate population.[103] Enlightenment philosophers turned to a few of their scientific predecessors – Galileo, Kepler, Boyle, and Newton principally – as the guides to every physical and social field of the day.[104][105] The 18th century saw significant advancements in the practice of medicine[106] and physics;[107] the development of biological taxonomy by Carl Linnaeus;[108] a new understanding of magnetism and electricity;[109] and the maturation of chemistry as a discipline.[110] Ideas on human nature, society, and economics evolved during the Enlightenment. Hume and other Scottish Enlightenment thinkers developed A Treatise of Human Nature, which was expressed historically in works by authors including James Burnett, Adam Ferguson, John Millar and William Robertson, all of whom merged a scientific study of how humans behaved in ancient and primitive cultures with a strong awareness of the determining forces of modernity.[111] Modern sociology largely originated from this movement.[112] In 1776, Adam Smith published The Wealth of Nations, which is often considered the first work on modern economics.[113] 19th century Main article: 19th century in science Sketch of a map with captions The first diagram of an evolutionary tree made by Charles Darwin in 1837 During the nineteenth century, many distinguishing characteristics of contemporary modern science began to take shape. These included the transformation of the life and physical sciences, frequent use of precision instruments, emergence of terms such as "biologist", "physicist", "scientist", increased professionalization of those studying nature, scientists gained cultural authority over many dimensions of society, industrialization of numerous countries, thriving of popular science writings and emergence of science journals.[114] During the late 19th century, psychology emerged as a separate discipline from philosophy when Wilhelm Wundt founded the first laboratory for psychological research in 1879.[115] During the mid-19th century, Charles Darwin and Alfred Russel Wallace independently proposed the theory of evolution by natural selection in 1858, which explained how different plants and animals originated and evolved. Their theory was set out in detail in Darwin's book On the Origin of Species, published in 1859.[116] Separately, Gregor Mendel presented his paper, "Experiments on Plant Hybridization" in 1865,[117] which outlined the principles of biological inheritance, serving as the basis for modern genetics.[118] Early in the 19th century, John Dalton suggested the modern atomic theory, based on Democritus's original idea of indivisible particles called atoms.[119] The laws of conservation of energy, conservation of momentum and conservation of mass suggested a highly stable universe where there could be little loss of resources. However, with the advent of the steam engine and the industrial revolution there was an increased understanding that not all forms of energy have the same energy qualities, the ease of conversion to useful work or to another form of energy.[120] This realization led to the development of the laws of thermodynamics, in which the free energy of the universe is seen as constantly declining: the entropy of a closed universe increases over time.[b] The electromagnetic theory was established in the 19th century by the works of Hans Christian Ørsted, André-Marie Ampère, Michael Faraday, James Clerk Maxwell, Oliver Heaviside, and Heinrich Hertz. The new theory raised questions that could not easily be answered using Newton's framework. The discovery of X-rays inspired the discovery of radioactivity by Henri Becquerel and Marie Curie in 1896,[123] Marie Curie then became the first person to win two Nobel prizes.[124] In the next year came the discovery of the first subatomic particle, the electron.[125] 20th century Main article: 20th century in science Graph showing lower ozone concentration at the South Pole A computer graph of the ozone hole made in 1987 using data from a space telescope In the first half of the century, the development of antibiotics and artificial fertilizers improved human living standards globally.[126][127] Harmful environmental issues such as ozone depletion, ocean acidification, eutrophication and climate change came to the public's attention and caused the onset of environmental studies.[128] During this period, scientific experimentation became increasingly larger in scale and funding.[129] The extensive technological innovation stimulated by World War I, World War II, and the Cold War led to competitions between global powers, such as the Space Race and nuclear arms race.[130][131] Substantial international collaborations were also made, despite armed conflicts.[132] In the late 20th century, active recruitment of women and elimination of sex discrimination greatly increased the number of women scientists, but large gender disparities remained in some fields.[133] The discovery of the cosmic microwave background in 1964[134] led to a rejection of the steady-state model of the universe in favor of the Big Bang theory of Georges Lemaître.[135] The century saw fundamental changes within science disciplines. Evolution became a unified theory in the early 20th-century when the modern synthesis reconciled Darwinian evolution with classical genetics.[136] Albert Einstein's theory of relativity and the development of quantum mechanics complement classical mechanics to describe physics in extreme length, time and gravity.[137][138] Widespread use of integrated circuits in the last quarter of the 20th century combined with communications satellites led to a revolution in information technology and the rise of the global internet and mobile computing, including smartphones. The need for mass systematization of long, intertwined causal chains and large amounts of data led to the rise of the fields of systems theory and computer-assisted scientific modeling.[139] 21st century Main article: 21st century § Science and technology Four predicted image of M87* black hole made by separate teams in the Event Horizon Telescope collaboration. The Human Genome Project was completed in 2003 by identifying and mapping all of the genes of the human genome.[140] The first induced pluripotent human stem cells were made in 2006, allowing adult cells to be transformed into stem cells and turn to any cell type found in the body.[141] With the affirmation of the Higgs boson discovery in 2013, the last particle predicted by the Standard Model of particle physics was found.[142] In 2015, gravitational waves, predicted by general relativity a century before, were first observed.[143][144] In 2019, the international collaboration Event Horizon Telescope presented the first direct image of a black hole's accretion disk.[145] Branches Main article: Branches of science Modern science is commonly divided into three major branches: natural science, social science, and formal science.[3] Each of these branches comprises various specialized yet overlapping scientific disciplines that often possess their own nomenclature and expertise.[146] Both natural and social sciences are empirical sciences,[147] as their knowledge is based on empirical observations and is capable of being tested for its validity by other researchers working under the same conditions.[148] Natural science Natural science is the study of the physical world. It can be divided into two main branches: life science and physical science. These two branches may be further divided into more specialized disciplines. For example, physical science can be subdivided into physics, chemistry, astronomy, and earth science. Modern natural science is the successor to the natural philosophy that began in Ancient Greece. Galileo, Descartes, Bacon, and Newton debated the benefits of using approaches which were more mathematical and more experimental in a methodical way. Still, philosophical perspectives, conjectures, and presuppositions, often overlooked, remain necessary in natural science.[149] Systematic data collection, including discovery science, succeeded natural history, which emerged in the 16th century by describing and classifying plants, animals, minerals, and so on.[150] Today, "natural history" suggests observational descriptions aimed at popular audiences.[151] Social science Two curve crossing over at a point, forming a X shape Supply and demand curve in economics, crossing over at the optimal equilibrium Social science is the study of human behavior and functioning of societies.[4][5] It has many disciplines that include, but are not limited to anthropology, economics, history, human geography, political science, psychology, and sociology.[4] In the social sciences, there are many competing theoretical perspectives, many of which are extended through competing research programs such as the functionalists, conflict theorists, and interactionists in sociology.[4] Due to the limitations of conducting controlled experiments involving large groups of individuals or complex situations, social scientists may adopt other research methods such as the historical method, case studies, and cross-cultural studies. Moreover, if quantitative information is available, social scientists may rely on statistical approaches to better understand social relationships and processes.[4] Formal science Formal science is an area of study that generates knowledge using formal systems.[152][6][7] A formal system is an abstract structure used for inferring theorems from axioms according to a set of rules.[153] It includes mathematics,[154][155] systems theory, and theoretical computer science. The formal sciences share similarities with the other two branches by relying on objective, careful, and systematic study of an area of knowledge. They are, however, different from the empirical sciences as they rely exclusively on deductive reasoning, without the need for empirical evidence, to verify their abstract concepts.[11][156][148] The formal sciences are therefore a priori disciplines and because of this, there is disagreement on whether they constitute a science.[8][157] Nevertheless, the formal sciences play an important role in the empirical sciences. Calculus, for example, was initially invented to understand motion in physics.[158] Natural and social sciences that rely heavily on mathematical applications include mathematical physics,[159] chemistry,[160] biology,[161] finance,[162] and economics.[163] Applied science Applied science is the use of the scientific method and knowledge to attain practical goals and includes a broad range of disciplines such as engineering and medicine.[164][14] Engineering is the use of scientific principles to invent, design and build machines, structures and technologies.[165] Science may contribute to the development of new technologies.[166] Medicine is the practice of caring for patients by maintaining and restoring health through the prevention, diagnosis, and treatment of injury or disease.[167][168] The applied sciences are often contrasted with the basic sciences, which are focused on advancing scientific theories and laws that explain and predict events in the natural world.[169][170] Computational science applies computing power to simulate real-world situations, enabling a better understanding of scientific problems than formal mathematics alone can achieve. The use of machine learning and artificial intelligence is becoming a central feature of computational contributions to science for example in agent-based computational economics, random forests, topic modeling and various forms of prediction. However, machines alone rarely advance knowledge as they require human guidance and capacity to reason; and they can introduce bias against certain social groups or sometimes underperform against humans.[171][172] Interdisciplinary science Interdisciplinary science involves the combination of two or more disciplines into one,[173] such as bioinformatics, a combination of biology and computer science[174] or cognitive sciences. The concept has existed since the ancient Greek and it became popular again in the 20th century.[175] Scientific research Scientific research can be labeled as either basic or applied research. Basic research is the search for knowledge and applied research is the search for solutions to practical problems using this knowledge. Most understanding comes from basic research, though sometimes applied research targets specific practical problems. This leads to technological advances that were not previously imaginable.[176] Scientific method 6 steps of the scientific method in a loop A diagram variant of scientific method represented as an ongoing process[broken anchor] Scientific research involves using the scientific method, which seeks to objectively explain the events of nature in a reproducible way.[177] Scientists usually take for granted a set of basic assumptions that are needed to justify the scientific method: there is an objective reality shared by all rational observers; this objective reality is governed by natural laws; these laws were discovered by means of systematic observation and experimentation.[2] Mathematics is essential in the formation of hypotheses, theories, and laws, because it is used extensively in quantitative modeling, observing, and collecting measurements.[178] Statistics is used to summarize and analyze data, which allows scientists to assess the reliability of experimental results.[179] In the scientific method, an explanatory thought experiment or hypothesis is put forward as an explanation using parsimony principles and is expected to seek consilience – fitting with other accepted facts related to an observation or scientific question.[180] This tentative explanation is used to make falsifiable predictions, which are typically posted before being tested by experimentation. Disproof of a prediction is evidence of progress.[177]: 4–5 [181] Experimentation is especially important in science to help establish causal relationships to avoid the correlation fallacy, though in some sciences such as astronomy or geology, a predicted observation might be more appropriate.[182] When a hypothesis proves unsatisfactory, it is modified or discarded.[183] If the hypothesis survived testing, it may become adopted into the framework of a scientific theory, a validly reasoned, self-consistent model or framework for describing the behavior of certain natural events. A theory typically describes the behavior of much broader sets of observations than a hypothesis; commonly, a large number of hypotheses can be logically bound together by a single theory. Thus a theory is a hypothesis explaining various other hypotheses. In that vein, theories are formulated according to most of the same scientific principles as hypotheses. Scientists may generate a model, an attempt to describe or depict an observation in terms of a logical, physical or mathematical representation and to generate new hypotheses that can be tested by experimentation.[184] While performing experiments to test hypotheses, scientists may have a preference for one outcome over another.[185][186] Eliminating the bias can be achieved by transparency, careful experimental design, and a thorough peer review process of the experimental results and conclusions.[187][188] After the results of an experiment are announced or published, it is normal practice for independent researchers to double-check how the research was performed, and to follow up by performing similar experiments to determine how dependable the results might be.[189] Taken in its entirety, the scientific method allows for highly creative problem solving while minimizing the effects of subjective and confirmation bias.[190] Intersubjective verifiability, the ability to reach a consensus and reproduce results, is fundamental to the creation of all scientific knowledge.[191] Scientific literature Main articles: Scientific literature and Lists of important publications in science Decorated "NATURE" as title, with scientific text below Cover of the first issue of Nature, November 4, 1869 Scientific research is published in a range of literature.[192] Scientific journals communicate and document the results of research carried out in universities and various other research institutions, serving as an archival record of science. The first scientific journals, Journal des sçavans followed by Philosophical Transactions, began publication in 1665. Since that time the total number of active periodicals has steadily increased. In 1981, one estimate for the number of scientific and technical journals in publication was 11,500.[193] Most scientific journals cover a single scientific field and publish the research within that field; the research is normally expressed in the form of a scientific paper. Science has become so pervasive in modern societies that it is considered necessary to communicate the achievements, news, and ambitions of scientists to a wider population.[194] Challenges The replication crisis is an ongoing methodological crisis that affects parts of the social and life sciences. In subsequent investigations, the results of many scientific studies are proven to be unrepeatable.[195] The crisis has long-standing roots; the phrase was coined in the early 2010s[196] as part of a growing awareness of the problem. The replication crisis represents an important body of research in metascience, which aims to improve the quality of all scientific research while reducing waste.[197] An area of study or speculation that masquerades as science in an attempt to claim a legitimacy that it would not otherwise be able to achieve is sometimes referred to as pseudoscience, fringe science, or junk science.[198][199] Physicist Richard Feynman coined the term "cargo cult science" for cases in which researchers believe and at a glance looks like they are doing science, but lack the honesty allowing their results to be rigorously evaluated.[200] Various types of commercial advertising, ranging from hype to fraud, may fall into these categories. Science has been described as "the most important tool" for separating valid claims from invalid ones.[201] There can also be an element of political or ideological bias on all sides of scientific debates. Sometimes, research may be characterized as "bad science," research that may be well-intended but is incorrect, obsolete, incomplete, or over-simplified expositions of scientific ideas. The term "scientific misconduct" refers to situations such as where researchers have intentionally misrepresented their published data or have purposely given credit for a discovery to the wrong person.[202] Philosophy of science Depiction of epicycles, where a planet orbit is going around in a bigger orbit For Kuhn, the addition of epicycles in Ptolemaic astronomy was "normal science" within a paradigm, whereas the Copernican Revolution was a paradigm shift There are different schools of thought in the philosophy of science. The most popular position is empiricism, which holds that knowledge is created by a process involving observation; scientific theories generalize observations.[203] Empiricism generally encompasses inductivism, a position that explains how general theories can be made from the finite amount of empirical evidence available. Many versions of empiricism exist, with the predominant ones being Bayesianism and the hypothetico-deductive method.[204][203] Empiricism has stood in contrast to rationalism, the position originally associated with Descartes, which holds that knowledge is created by the human intellect, not by observation.[205] Critical rationalism is a contrasting 20th-century approach to science, first defined by Austrian-British philosopher Karl Popper. Popper rejected the way that empiricism describes the connection between theory and observation. He claimed that theories are not generated by observation, but that observation is made in the light of theories: that the only way theory A can be affected by observation is after theory A were to conflict with observation, but theory B were to survive the observation.[206] Popper proposed replacing verifiability with falsifiability as the landmark of scientific theories, replacing induction with falsification as the empirical method.[206] Popper further claimed that there is actually only one universal method, not specific to science: the negative method of criticism, trial and error,[207] covering all products of the human mind, including science, mathematics, philosophy, and art.[208] Another approach, instrumentalism, emphasizes the utility of theories as instruments for explaining and predicting phenomena. It views scientific theories as black boxes with only their input (initial conditions) and output (predictions) being relevant. Consequences, theoretical entities, and logical structure are claimed to be something that should be ignored.[209] Close to instrumentalism is constructive empiricism, according to which the main criterion for the success of a scientific theory is whether what it says about observable entities is true.[210] Thomas Kuhn argued that the process of observation and evaluation takes place within a paradigm, a logically consistent "portrait" of the world that is consistent with observations made from its framing. He characterized normal science as the process of observation and "puzzle solving" which takes place within a paradigm, whereas revolutionary science occurs when one paradigm overtakes another in a paradigm shift.[211] Each paradigm has its own distinct questions, aims, and interpretations. The choice between paradigms involves setting two or more "portraits" against the world and deciding which likeness is most promising. A paradigm shift occurs when a significant number of observational anomalies arise in the old paradigm and a new paradigm makes sense of them. That is, the choice of a new paradigm is based on observations, even though those observations are made against the background of the old paradigm. For Kuhn, acceptance or rejection of a paradigm is a social process as much as a logical process. Kuhn's position, however, is not one of relativism.[212] Finally, another approach often cited in debates of scientific skepticism against controversial movements like "creation science" is methodological naturalism. Naturalists maintain that a difference should be made between natural and supernatural, and science should be restricted to natural explanations.[213] Methodological naturalism maintains that science requires strict adherence to empirical study and independent verification.[214] Scientific community The scientific community is a network of interacting scientists who conducts scientific research. The community consists of smaller groups working in scientific fields. By having peer review, through discussion and debate within journals and conferences, scientists maintain the quality of research methodology and objectivity when interpreting results.[215] Scientists Portrait of a middle-aged woman Marie Curie was the first person to be awarded two Nobel Prizes: Physics in 1903 and Chemistry in 1911[124] Scientists are individuals who conduct scientific research to advance knowledge in an area of interest.[216][217] In modern times, many professional scientists are trained in an academic setting and upon completion, attain an academic degree, with the highest degree being a doctorate such as a Doctor of Philosophy or PhD.[218] Many scientists pursue careers in various sectors of the economy such as academia, industry, government, and nonprofit organizations.[219][220][221] Scientists exhibit a strong curiosity about reality and a desire to apply scientific knowledge for the benefit of health, nations, the environment, or industries. Other motivations include recognition by their peers and prestige. In modern times, many scientists have advanced degrees in an area of science and pursue careers in various sectors of the economy such as academia, industry, government, and nonprofit environments.[222] [223][224] Science has historically been a male-dominated field, with notable exceptions. Women in science faced considerable discrimination in science, much as they did in other areas of male-dominated societies. For example, women were frequently being passed over for job opportunities and denied credit for their work.[225] The achievements of women in science have been attributed to the defiance of their traditional role as laborers within the domestic sphere.[226] Learned societies Picture of scientists in 200th anniversary of the Prussian Academy of Sciences, 1900 Learned societies for the communication and promotion of scientific thought and experimentation have existed since the Renaissance.[227] Many scientists belong to a learned society that promotes their respective scientific discipline, profession, or group of related disciplines.[228] Membership may either be open to all, require possession of scientific credentials, or conferred by election.[229] Most scientific societies are non-profit organizations,[230] and many are professional associations. Their activities typically include holding regular conferences for the presentation and discussion of new research results and publishing or sponsoring academic journals in their discipline. Some societies act as professional bodies, regulating the activities of their members in the public interest, or the collective interest of the membership. The professionalization of science, begun in the 19th century, was partly enabled by the creation of national distinguished academies of sciences such as the Italian Accademia dei Lincei in 1603,[231] the British Royal Society in 1660,[232] the French Academy of Sciences in 1666,[233] the American National Academy of Sciences in 1863,[234] the German Kaiser Wilhelm Society in 1911,[235] and the Chinese Academy of Sciences in 1949.[236] International scientific organizations, such as the International Science Council, are devoted to international cooperation for science advancement.[237] Awards Science awards are usually given to individuals or organizations that have made significant contributions to a discipline. They are often given by prestigious institutions, thus it is considered a great honor for a scientist receiving them. Since the early Renaissance, scientists are often awarded medals, money, and titles. The Nobel Prize, a widely regarded prestigious award, is awarded annually to those who have achieved scientific advances in the fields of medicine, physics, and chemistry.[238] Society "Science and society" redirects here. Not to be confused with Science & Society or Sociology of scientific knowledge. Funding and policies see caption Budget of NASA as percentage of United States federal budget, peaking at 4.4% in 1966 and slowly declining since Scientific research is often funded through a competitive process in which potential research projects are evaluated and only the most promising receive funding. Such processes, which are run by government, corporations, or foundations, allocate scarce funds. Total research funding in most developed countries is between 1.5% and 3% of GDP.[239] In the OECD, around two-thirds of research and development in scientific and technical fields is carried out by industry, and 20% and 10% respectively by universities and government. The government funding proportion in certain fields is higher, and it dominates research in social science and humanities. In the lesser-developed nations, government provides the bulk of the funds for their basic scientific research.[240] Many governments have dedicated agencies to support scientific research, such as the National Science Foundation in the United States,[241] the National Scientific and Technical Research Council in Argentina,[242] Commonwealth Scientific and Industrial Research Organization in Australia,[243] National Centre for Scientific Research in France,[244] the Max Planck Society in Germany,[245] and National Research Council in Spain.[246] In commercial research and development, all but the most research-oriented corporations focus more heavily on near-term commercialization possibilities rather than research driven by curiosity.[247] Science policy is concerned with policies that affect the conduct of the scientific enterprise, including research funding, often in pursuance of other national policy goals such as technological innovation to promote commercial product development, weapons development, health care, and environmental monitoring. Science policy sometimes refers to the act of applying scientific knowledge and consensus to the development of public policies. In accordance with public policy being concerned about the well-being of its citizens, science policy's goal is to consider how science and technology can best serve the public.[248] Public policy can directly affect the funding of capital equipment and intellectual infrastructure for industrial research by providing tax incentives to those organizations that fund research.[194] Education and awareness Main articles: Public awareness of science and Science journalism Dinosaur exhibit in the Houston Museum of Natural Science Science education for the general public is embedded in the school curriculum, and is supplemented by online pedagogical content (for example, YouTube and Khan Academy), museums, and science magazines and blogs. Scientific literacy is chiefly concerned with an understanding of the scientific method, units and methods of measurement, empiricism, a basic understanding of statistics (correlations, qualitative versus quantitative observations, aggregate statistics), as well as a basic understanding of core scientific fields, such as physics, chemistry, biology, ecology, geology and computation. As a student advances into higher stages of formal education, the curriculum becomes more in depth. Traditional subjects usually included in the curriculum are natural and formal sciences, although recent movements include social and applied science as well.[249] The mass media face pressures that can prevent them from accurately depicting competing scientific claims in terms of their credibility within the scientific community as a whole. Determining how much weight to give different sides in a scientific debate may require considerable expertise regarding the matter.[250] Few journalists have real scientific knowledge, and even beat reporters who are knowledgeable about certain scientific issues may be ignorant about other scientific issues that they are suddenly asked to cover.[251][252] Science magazines such as New Scientist, Science & Vie, and Scientific American cater to the needs of a much wider readership and provide a non-technical summary of popular areas of research, including notable discoveries and advances in certain fields of research.[253] Science fiction genre, primarily speculative fiction, can transmit the ideas and methods of science to the general public.[254] Recent efforts to intensify or develop links between science and non-scientific disciplines, such as literature or poetry, include the Creative Writing Science resource developed through the Royal Literary Fund.[255] Anti-science attitudes Main article: Antiscience While the scientific method is broadly accepted in the scientific community, some fractions of society reject certain scientific positions or are skeptical about science. Examples are the common notion that COVID-19 is not a major health threat to the US (held by 39% of Americans in August 2021)[256] or the belief that climate change is not a major threat to the US (also held by 40% of Americans, in late 2019 and early 2020).[257] Psychologists have pointed to four factors driving rejection of scientific results:[258] Scientific authorities are sometimes seen as inexpert, untrustworthy, or biased. Some marginalized social groups hold anti-science attitudes, in part because these groups have often been exploited in unethical experiments.[259] Messages from scientists may contradict deeply-held existing beliefs or morals. The delivery of a scientific message may not be appropriately targeted to a recipient's learning style. Anti-science attitudes seem to be often caused by fear of rejection in social groups. For instance, climate change is perceived as a threat by only 22% of Americans on the right side of the political spectrum, but by 85% on the left.[260] That is, if someone on the left would not consider climate change as a threat, this person may face contempt and be rejected in that social group. In fact, people may rather deny a scientifically accepted fact than lose or jeopardize their social status.[261] Politics Result in bar graph of two questions ("Is global warming occurring?" and "Are oil/gas companies responsible?"), showing large discrepancies between American Democrats and Republicans Public opinion on global warming in the United States by political party[262] Attitudes towards science are often determined by political opinions and goals. Government, business and advocacy groups have been known to use legal and economic pressure to influence scientific researchers. Many factors can act as facets of the politicization of science such as anti-intellectualism, perceived threats to religious beliefs, and fear for business interests.[263] Politicization of science is usually accomplished when scientific information is presented in a way that emphasizes the uncertainty associated with the scientific evidence.[264] Tactics such as shifting conversation, failing to acknowledge facts, and capitalizing on doubt of scientific consensus have been used to gain more attention for views that have been undermined by scientific evidence.[265] Examples of issues that have involved the politicization of science include the global warming controversy, health effects of pesticides, and health effects of tobacco.[265][266] See also Criticism of science List of scientific occupations List of years in science Notes Ibn al-Haytham's Book of Optics Book I, [6.54]. pages 372 and 408 disputed Claudius Ptolemy's extramission theory of vision; "Hence, the extramission of [visual] rays is superfluous and useless". —A.Mark Smith's translation of the Latin version of Ibn al-Haytham.[84]: Book I, [6.54]. pp. 372, 408  Whether the universe is closed or open, or the shape of the universe, is an open question. The 2nd law of thermodynamics,[120]: 9 [121] and the 3rd law of thermodynamics[122] imply the heat death of the universe if the universe is a closed system, but not necessarily for an expanding universe. References Wilson, E.O. (1999). "The natural sciences". Consilience: The Unity of Knowledge (Reprint ed.). New York: Vintage. pp. 49–71. ISBN 978-0-679-76867-8. Heilbron, J.L.; et al. (2003). "Preface". The Oxford Companion to the History of Modern Science. New York: Oxford University Press. pp. vii–x. ISBN 978-0-19-511229-0. ...modern science is a discovery as well as an invention. It was a discovery that nature generally acts regularly enough to be described by laws and even by mathematics; and required invention to devise the techniques, abstractions, apparatus, and organization for exhibiting the regularities and securing their law-like descriptions. Cohen, Eliel (2021). "The boundary lens: theorising academic actitity". The University and its Boundaries: Thriving or Surviving in the 21st Century. New York: Routledge. pp. 14–41. ISBN 978-0-367-56298-4. Archived from the original on May 5, 2021. Retrieved May 4, 2021. Colander, David C.; Hunt, Elgin F. (2019). "Social science and its methods". Social Science: An Introduction to the Study of Society (17th ed.). New York, NY: Routledge. pp. 1–22. Nisbet, Robert A.; Greenfeld, Liah (October 16, 2020). "Social Science". Encyclopedia Britannica. Encyclopædia Britannica, Inc. Archived from the original on February 2, 2022. Retrieved May 9, 2021. Löwe, Benedikt (2002). "The formal sciences: their scope, their foundations, and their unity". Synthese. 133 (1/2): 5–11. doi:10.1023/A:1020887832028. S2CID 9272212. Rucker, Rudy (2019). "Robots and souls". Infinity and the Mind: The Science and Philosophy of the Infinite (Reprint ed.). Princeton, New Jersey: Princeton University Press. pp. 157–188. ISBN 978-0-691-19138-6. Archived from the original on February 26, 2021. Retrieved May 11, 2021. Bishop, Alan (1991). "Environmental activities and mathematical culture". Mathematical Enculturation: A Cultural Perspective on Mathematics Education. Norwell, Massachusetts: Kluwer Academic Publishers. pp. 20–59. ISBN 978-0-7923-1270-3. Archived from the original on December 25, 2020. Retrieved March 24, 2018. Nickles, Thomas (2013). "The Problem of Demarcation". Philosophy of Pseudoscience: Reconsidering the Demarcation Problem. Chicago: The University of Chicago Press. p. 104. Bunge, Mario (1998). "The Scientific Approach". Philosophy of Science. Vol. 1, From Problem to Theory (revised ed.). New York: Routledge. pp. 3–50. ISBN 978-0-7658-0413-6. Fetzer, James H. (2013). "Computer reliability and public policy: Limits of knowledge of computer-based systems". Computers and Cognition: Why Minds are not Machines. Newcastle, United Kingdom: Kluwer Academic Publishers. pp. 271–308. ISBN 978-1-4438-1946-6. Fischer, M.R.; Fabry, G (2014). "Thinking and acting scientifically: Indispensable basis of medical education". GMS Zeitschrift für Medizinische Ausbildung. 31 (2): Doc24. doi:10.3205/zma000916. PMC 4027809. PMID 24872859. Sinclair, Marius (1993). "On the Differences between the Engineering and Scientific Methods". The International Journal of Engineering Education. Archived from the original on November 15, 2017. Retrieved September 7, 2018. Bunge, M (1966). "Technology as Applied Science". In Rapp, F. (ed.). Contributions to a Philosophy of Technology. Dordrecht, Netherlands: Springer. pp. 19–39. doi:10.1007/978-94-010-2182-1_2. ISBN 978-94-010-2184-5. S2CID 110332727. Lindberg, David C. (2007). The beginnings of Western science: the European Scientific tradition in philosophical, religious, and institutional context (2nd ed.). University of Chicago Press. ISBN 978-0226482057. Grant, Edward (2007). "Ancient Egypt to Plato". A History of Natural Philosophy: From the Ancient World to the Nineteenth Century. New York: Cambridge University Press. pp. 1–26. ISBN 978-0-521-68957-1. Building Bridges Among the BRICs, p. 125, Robert Crane, Springer, 2014 Keay, John (2000). India: A history. Atlantic Monthly Press. p. 132. ISBN 978-0-87113-800-2. The great era of all that is deemed classical in Indian literature, art and science was now dawning. It was this crescendo of creativity and scholarship, as much as ... political achievements of the Guptas, which would make their age so golden. Lindberg, David C. (2007). "Islamic science". The beginnings of Western science: the European Scientific tradition in philosophical, religious, and institutional context (2nd ed.). Chicago: University of Chicago Press. pp. 163–92. ISBN 978-0-226-48205-7. Lindberg, David C. (2007). "The revival of learning in the West". The beginnings of Western science: the European Scientific tradition in philosophical, religious, and institutional context (2nd ed.). Chicago: University of Chicago Press. pp. 193–224. ISBN 978-0-226-48205-7. Lindberg, David C. (2007). "The recovery and assimilation of Greek and Islamic science". The beginnings of Western science: the European Scientific tradition in philosophical, religious, and institutional context (2nd ed.). Chicago: University of Chicago Press. pp. 225–53. ISBN 978-0-226-48205-7. Sease, Virginia; Schmidt-Brabant, Manfrid. Thinkers, Saints, Heretics: Spiritual Paths of the Middle Ages. 2007. Pages 80-81. Retrieved 6 Oct. 2023 Principe, Lawrence M. (2011). "Introduction". Scientific Revolution: A Very Short Introduction. New York: Oxford University Press. pp. 1–3. ISBN 978-0-19-956741-6. Lindberg, David C. (2007). "The legacy of ancient and medieval science". The beginnings of Western science: the European Scientific tradition in philosophical, religious, and institutional context (2nd ed.). Chicago: University of Chicago Press. pp. 357–368. ISBN 978-0-226-48205-7. Grant, Edward (2007). "Transformation of medieval natural philosophy from the early period modern period to the end of the nineteenth century". A History of Natural Philosophy: From the Ancient World to the Nineteenth Century. New York: Cambridge University Press. pp. 274–322. ISBN 978-0-521-68957-1. Cahan, David, ed. (2003). From Natural Philosophy to the Sciences: Writing the History of Nineteenth-Century Science. Chicago: University of Chicago Press. ISBN 978-0-226-08928-7. Lightman, Bernard (2011). "13. Science and the Public". In Shank, Michael; Numbers, Ronald; Harrison, Peter (eds.). Wrestling with Nature: From Omens to Science. Chicago: University of Chicago Press. p. 367. ISBN 978-0-226-31783-0. Harrison, Peter (2015). The Territories of Science and Religion. Chicago: University of Chicago Press. pp. 164–165. ISBN 978-0-226-18451-7. The changing character of those engaged in scientific endeavors was matched by a new nomenclature for their endeavors. The most conspicuous marker of this change was the replacement of "natural philosophy" by "natural science". In 1800 few had spoken of the "natural sciences" but by 1880, this expression had overtaken the traditional label "natural philosophy". The persistence of "natural philosophy" in the twentieth century is owing largely to historical references to a past practice (see figure 11). As should now be apparent, this was not simply the substitution of one term by another, but involved the jettisoning of a range of personal qualities relating to the conduct of philosophy and the living of the philosophical life. MacRitchie, Finlay (2011). "Introduction". Scientific Research as a Career. New York: Routledge. pp. 1–6. ISBN 978-1-4398-6965-9. Archived from the original on May 5, 2021. Retrieved May 5, 2021. Marder, Michael P. (2011). "Curiosity and research". Research Methods for Science. New York: Cambridge University Press. pp. 1–17. ISBN 978-0-521-14584-8. Archived from the original on May 5, 2021. Retrieved May 5, 2021. de Ridder, Jeroen (2020). "How many scientists does it take to have knowledge?". In McCain, Kevin; Kampourakis, Kostas (eds.). What is Scientific Knowledge? An Introduction to Contemporary Epistemology of Science. New York: Routledge. pp. 3–17. ISBN 978-1-138-57016-0. Archived from the original on May 5, 2021. Retrieved May 5, 2021. Lindberg, David C. (2007). "Islamic science". The beginnings of Western science: the European Scientific tradition in philosophical, religious, and institutional context (2nd ed.). Chicago: University of Chicago Press. pp. 163–192. ISBN 978-0-226-48205-7. Szycher, Michael (2016). "Establishing your dream team". Commercialization Secrets for Scientists and Engineers. New York: Routledge. pp. 159–176. ISBN 978-1-138-40741-1. Archived from the original on August 18, 2021. Retrieved May 5, 2021. "science". Merriam-Webster Online Dictionary. Merriam-Webster, Inc. Archived from the original on September 1, 2019. Retrieved October 16, 2011. Vaan, Michiel de (2008). "sciō". Etymological Dictionary of Latin and the other Italic Languages. Indo-European Etymological Dictionary. p. 545. ISBN 978-90-04-16797-1. Cahan, David (2003). From natural philosophy to the sciences : writing the history of nineteenth-century science. Chicago: University of Chicago Press. pp. 3–15. ISBN 0-226-08927-4. OCLC 51330464. Archived from the original on May 31, 2022. Retrieved May 31, 2022. Ross, Sydney (1962). "Scientist: The story of a word". Annals of Science. 18 (2): 65–85. doi:10.1080/00033796200202722. "scientist". Oxford English Dictionary (Online ed.). Oxford University Press. (Subscription or participating institution membership required.) Carruthers, Peter (May 2, 2002). Carruthers, Peter; Stich, Stephen; Siegal, Michael (eds.). "The roots of scientific reasoning: infancy, modularity and the art of tracking". The Cognitive Basis of Science. Cambridge University Press. pp. 73–96. doi:10.1017/cbo9780511613517.005. ISBN 978-0-521-81229-0. Lombard, Marlize; Gärdenfors, Peter (2017). "Tracking the Evolution of Causal Cognition in Humans". Journal of Anthropological Sciences. 95 (95): 219–234. doi:10.4436/JASS.95006. ISSN 1827-4765. PMID 28489015. Graeber, David; Wengrow, David (2021). The Dawn of Everything. p. 248. Budd, Paul; Taylor, Timothy (1995). "The Faerie Smith Meets the Bronze Industry: Magic Versus Science in the Interpretation of Prehistoric Metal-Making". World Archaeology. 27 (1): 133–143. doi:10.1080/00438243.1995.9980297. JSTOR 124782. Tuomela, Raimo (1987). "Science, Protoscience, and Pseudoscience". In Pitt, J.C.; Pera, M. (eds.). Rational Changes in Science. Boston Studies in the Philosophy of Science. Vol. 98. Dordrecht: Springer. pp. 83–101. doi:10.1007/978-94-009-3779-6_4. ISBN 978-94-010-8181-8. Smith, Pamela H. (2009). "Science on the Move: Recent Trends in the History of Early Modern Science". Renaissance Quarterly. 62 (2): 345–375. doi:10.1086/599864. PMID 19750597. S2CID 43643053. Fleck, Robert (March 2021). "Fundamental Themes in Physics from the History of Art". Physics in Perspective. 23 (1): 25–48. Bibcode:2021PhP....23...25F. doi:10.1007/s00016-020-00269-7. ISSN 1422-6944. S2CID 253597172. Scott, Colin (2011). "Science for the West, Myth for the Rest?". In Harding, Sandra (ed.). The Postcolonial Science and Technology Studies Reader. Durham: Duke University Press. p. 175. doi:10.2307/j.ctv11g96cc.16. ISBN 978-0-8223-4936-5. OCLC 700406626. Dear, Peter (2012). "Historiography of Not-So-Recent Science". History of Science. 50 (2): 197–211. doi:10.1177/007327531205000203. S2CID 141599452. Rochberg, Francesca (2011). "Ch.1 Natural Knowledge in Ancient Mesopotamia". In Shank, Michael; Numbers, Ronald; Harrison, Peter (eds.). Wrestling with Nature: From Omens to Science. Chicago: University of Chicago Press. p. 9. ISBN 978-0-226-31783-0. Krebs, Robert E. (2004). Groundbreaking Scientific Experiments, Inventions, and Discoveries of the Middle Ages and the Renaissance. Greenwood Publishing Group. p. 127. ISBN 978-0313324338. Erlich, Ḥaggai; Gershoni, Israel (2000). The Nile: Histories, Cultures, Myths. Lynne Rienner Publishers. pp. 80–81. ISBN 978-1-55587-672-2. Archived from the original on May 31, 2022. Retrieved January 9, 2020. The Nile occupied an important position in Egyptian culture; it influenced the development of mathematics, geography, and the calendar; Egyptian geometry advanced due to the practice of land measurement "because the overflow of the Nile caused the boundary of each person's land to disappear." "Telling Time in Ancient Egypt". The Met's Heilbrunn Timeline of Art History. Archived from the original on March 3, 2022. Retrieved May 27, 2022. McIntosh, Jane R. (2005). Ancient Mesopotamia: New Perspectives. Santa Barbara, California, Denver, Colorado, and Oxford, England: ABC-CLIO. pp. 273–76. ISBN 978-1-57607-966-9. Archived from the original on February 5, 2021. Retrieved October 20, 2020. Aaboe, Asger (May 2, 1974). "Scientific Astronomy in Antiquity". Philosophical Transactions of the Royal Society. 276 (1257): 21–42. Bibcode:1974RSPTA.276...21A. doi:10.1098/rsta.1974.0007. JSTOR 74272. S2CID 122508567. Biggs, R D. (2005). "Medicine, Surgery, and Public Health in Ancient Mesopotamia". Journal of Assyrian Academic Studies. 19 (1): 7–18. Lehoux, Daryn (2011). "2. Natural Knowledge in the Classical World". In Shank, Michael; Numbers, Ronald; Harrison, Peter (eds.). Wrestling with Nature: From Omens to Science. Chicago: University of Chicago Press. p. 39. ISBN 978-0-226-31783-0. An account of the pre-Socratic use of the concept of φύσις may be found in Naddaf, Gerard (2006). The Greek Concept of Nature. SUNY Press, and in Ducarme, Frédéric; Couvet, Denis (2020). "What does 'nature' mean?" (PDF). Palgrave Communications. 6 (14). Springer Nature. doi:10.1057/s41599-020-0390-y. The word φύσις, while first used in connection with a plant in Homer, occurs early in Greek philosophy, and in several senses. Generally, these senses match rather well the current senses in which the English word nature is used, as confirmed by Guthrie, W.K.C. Presocratic Tradition from Parmenides to Democritus (volume 2 of his History of Greek Philosophy), Cambridge UP, 1965. Strauss, Leo; Gildin, Hilail (1989). "Progress or Return? The Contemporary Crisis in Western Education". An Introduction to Political Philosophy: Ten Essays by Leo Strauss. Wayne State University Press (published August 1, 1989). p. 209. ISBN 978-0814319024. Archived from the original on May 31, 2022. Retrieved May 30, 2022. O'Grady, Patricia F. (2016). Thales of Miletus: The Beginnings of Western Science and Philosophy. New York City, New York and London, England: Routledge. p. 245. ISBN 978-0-7546-0533-1. Archived from the original on March 31, 2021. Retrieved October 20, 2020. Burkert, Walter (June 1, 1972). Lore and Science in Ancient Pythagoreanism. Cambridge, Massachusetts: Harvard University Press. ISBN 978-0-674-53918-1. Archived from the original on January 29, 2018. Pullman, Bernard (1998). The Atom in the History of Human Thought. Oxford University Press. pp. 31–33. Bibcode:1998ahht.book.....P. ISBN 978-0-19-515040-7. Archived from the original on February 5, 2021. Retrieved October 20, 2020. Cohen, Henri; Lefebvre, Claire, eds. (2017). Handbook of Categorization in Cognitive Science (2nd ed.). Amsterdam, the Netherlands: Elsevier. p. 427. ISBN 978-0-08-101107-2. Archived from the original on February 5, 2021. Retrieved October 20, 2020. Lucretius (fl. 1st c. BCE) De rerum natura Margotta, Roberto (1968). The Story of Medicine. New York City, New York: Golden Press. Archived from the original on February 5, 2021. Retrieved November 18, 2020. Touwaide, Alain (2005). Glick, Thomas F.; Livesey, Steven; Wallis, Faith (eds.). Medieval Science, Technology, and Medicine: An Encyclopedia. New York City, New York and London, England: Routledge. p. 224. ISBN 978-0-415-96930-7. Archived from the original on February 6, 2021. Retrieved October 20, 2020. Leff, Samuel; Leff, Vera (1956). From Witchcraft to World Health. London, England: Macmillan. Archived from the original on February 5, 2021. Retrieved August 23, 2020. "Plato, Apology". p. 17. Archived from the original on January 29, 2018. Retrieved November 1, 2017. "Plato, Apology". p. 27. Archived from the original on January 29, 2018. Retrieved November 1, 2017. Aristotle. Nicomachean Ethics (H. Rackham ed.). 1139b. Archived from the original on March 17, 2012. Retrieved September 22, 2010. McClellan III, James E.; Dorn, Harold (2015). Science and Technology in World History: An Introduction. Baltimore, Maryland: Johns Hopkins University Press. pp. 99–100. ISBN 978-1-4214-1776-9. Archived from the original on February 6, 2021. Retrieved October 20, 2020. Graßhoff, Gerd (1990). The History of Ptolemy's Star Catalogue. Studies in the History of Mathematics and Physical Sciences. Vol. 14. New York, NY: Springer New York. doi:10.1007/978-1-4612-4468-4. ISBN 978-1-4612-8788-9. Archived from the original on May 30, 2022. Retrieved May 27, 2022. Hoffmann, Susanne M. (2017). Hipparchs Himmelsglobus (in German). Wiesbaden: Springer Fachmedien Wiesbaden. Bibcode:2017hihi.book.....H. doi:10.1007/978-3-658-18683-8. ISBN 978-3-658-18682-1. Archived from the original on May 30, 2022. Retrieved May 27, 2022. Edwards, C.H. Jr. (1979). The Historical Development of the Calculus. New York City, New York: Springer-Verlag. p. 75. ISBN 978-0-387-94313-8. Archived from the original on February 5, 2021. Retrieved October 20, 2020. Lawson, Russell M. (2004). Science in the Ancient World: An Encyclopedia. Santa Barbara, California: ABC-CLIO. pp. 190–91. ISBN 978-1-85109-539-1. Archived from the original on February 5, 2021. Retrieved October 20, 2020. Murphy, Trevor Morgan (2004). Pliny the Elder's Natural History: The Empire in the Encyclopedia. Oxford, England: Oxford University Press. p. 1. ISBN 978-0-19-926288-5. Archived from the original on February 6, 2021. Retrieved October 20, 2020. Doody, Aude (2010). Pliny's Encyclopedia: The Reception of the Natural History. Cambridge, England: Cambridge University Press. p. 1. ISBN 978-1-139-48453-4. Archived from the original on March 31, 2021. Retrieved October 20, 2020. Conner, Clifford D. (2005). A People's History of Science: Miners, Midwives, and "Low Mechanicks". New York: Nation Books. pp. 72–74. ISBN 1-56025-748-2. OCLC 62164511. Grant, Edward (1996). The Foundations of Modern Science in the Middle Ages: Their Religious, Institutional and Intellectual Contexts. Cambridge Studies in the History of Science. Cambridge University Press. pp. 7–17. ISBN 978-0-521-56762-6. Archived from the original on August 21, 2019. Retrieved November 9, 2018. Wildberg, Christian (May 1, 2018). Zalta, Edward N. (ed.). The Stanford Encyclopedia of Philosophy. Metaphysics Research Lab, Stanford University. Archived from the original on August 22, 2019. Retrieved May 1, 2018 – via Stanford Encyclopedia of Philosophy. Falcon, Andrea (2019). "Aristotle on Causality". In Zalta, Edward (ed.). Stanford Encyclopedia of Philosophy (Spring 2019 ed.). Metaphysics Research Lab, Stanford University. Archived from the original on October 9, 2020. Retrieved October 3, 2020. Grant, Edward (2007). "Islam and the eastward shift of Aristotelian natural philosophy". A History of Natural Philosophy: From the Ancient World to the Nineteenth Century. Cambridge University Press. pp. 62–67. ISBN 978-0-521-68957-1. Fisher, W.B. (William Bayne) (1968–1991). The Cambridge history of Iran. Cambridge: University Press. ISBN 978-0-521-20093-6. OCLC 745412. "Bayt al-Hikmah". Encyclopædia Britannica. Archived from the original on November 4, 2016. Retrieved November 3, 2016. Hossein Nasr, Seyyed; Leaman, Oliver, eds. (2001). History of Islamic Philosophy. Routledge. pp. 165–167. ISBN 978-0415259347. Smith, A. Mark (2001). Alhacen's Theory of Visual Perception: A Critical Edition, with English Translation and Commentary, of the First Three Books of Alhacen's De Aspectibus, the Medieval Latin Version of Ibn al-Haytham's Kitāb al-Manāẓir, 2 vols. Transactions of the American Philosophical Society. Vol. 91. Philadelphia: American Philosophical Society. ISBN 978-0-87169-914-5. OCLC 47168716. Toomer, G.J. (1964). "Reviewed work: Ibn al-Haythams Weg zur Physik, Matthias Schramm". Isis. 55 (4): 463–65. doi:10.1086/349914. JSTOR 228328. See p. 464: "Schramm sums up [Ibn Al-Haytham's] achievement in the development of scientific method.", p. 465: "Schramm has demonstrated .. beyond any dispute that Ibn al-Haytham is a major figure in the Islamic scientific tradition, particularly in the creation of experimental techniques." p. 465: "only when the influence of Ibn al-Haytham and others on the mainstream of later medieval physical writings has been seriously investigated can Schramm's claim that Ibn al-Haytham was the true founder of modern physics be evaluated." Cohen, H. Floris (2010). "Greek nature knowledge transplanted: The Islamic world". How modern science came into the world. Four civilizations, one 17th-century breakthrough (2nd ed.). Amsterdam: Amsterdam University Press. pp. 99–156. ISBN 978-90-8964-239-4. Selin, Helaine, ed. (2006). Encyclopaedia of the History of Science, Technology, and Medicine in Non-Western Cultures. Springer. pp. 155–156. Bibcode:2008ehst.book.....S. ISBN 978-1-4020-4559-2. Russell, Josiah C. (1959). "Gratian, Irnerius, and the Early Schools of Bologna". The Mississippi Quarterly. 12 (4): 168–188. JSTOR 26473232. Archived from the original on May 27, 2022. Retrieved May 27, 2022 – via JSTOR. Perhaps even as early as 1088 (the date officially set for the founding of the University) "St. Albertus Magnus | German theologian, scientist, and philosopher". Archived from the original on October 28, 2017. Retrieved October 27, 2017. Numbers, Ronald (2009). Galileo Goes to Jail and Other Myths about Science and Religion. Harvard University Press. p. 45. ISBN 978-0-674-03327-6. Archived from the original on January 20, 2021. Retrieved March 27, 2018. Smith, A. Mark (1981). "Getting the Big Picture in Perspectivist Optics". Isis. 72 (4): 568–89. doi:10.1086/352843. JSTOR 231249. PMID 7040292. S2CID 27806323. Goldstein, Bernard R (2016). "Copernicus and the Origin of his Heliocentric System" (PDF). Journal for the History of Astronomy. 33 (3): 219–35. doi:10.1177/002182860203300301. S2CID 118351058. Archived from the original (PDF) on April 12, 2020. Retrieved April 12, 2020. Cohen, H. Floris (2010). "Greek nature knowledge transplanted and more: Renaissance Europe". How modern science came into the world. Four civilizations, one 17th-century breakthrough (2nd ed.). Amsterdam: Amsterdam University Press. pp. 99–156. ISBN 978-90-8964-239-4. Koestler, Arthur (1990) [1959]. The Sleepwalkers: A History of Man's Changing Vision of the Universe. London: Penguin Books. p. 1. ISBN 0-14-019246-8. van Helden, Al (1995). "Pope Urban VIII". The Galileo Project. Archived from the original on November 11, 2016. Retrieved November 3, 2016. Gingerich, Owen (1975). "Copernicus and the Impact of Printing". Vistas in Astronomy. 17 (1): 201–218. Bibcode:1975VA.....17..201G. doi:10.1016/0083-6656(75)90061-6. Zagorin, Perez (1998). Francis Bacon. Princeton: Princeton University Press. p. 84. ISBN 978-0-691-00966-7. Davis, Philip J.; Hersh, Reuben (1986). Descartes' Dream: The World According to Mathematics. Cambridge, MA: Harcourt Brace Jovanovich. Gribbin, John (2002). Science: A History 1543–2001. Allen Lane. p. 241. ISBN 978-0-7139-9503-9. Although it was just one of the many factors in the Enlightenment, the success of Newtonian physics in providing a mathematical description of an ordered world clearly played a big part in the flowering of this movement in the eighteenth century "Gottfried Leibniz – Biography". Maths History. Archived from the original on July 11, 2017. Retrieved March 2, 2021. Freudenthal, Gideon; McLaughlin, Peter (May 20, 2009). The Social and Economic Roots of the Scientific Revolution: Texts by Boris Hessen and Henryk Grossmann. Springer Science & Business Media. ISBN 978-1-4020-9604-4. Archived from the original on January 19, 2020. Retrieved July 25, 2018. Goddard Bergin, Thomas; Speake, Jennifer, eds. (1987). Encyclopedia of the Renaissance. Facts on File (published December 1, 1987). ISBN 978-0816013159. van Horn Melton, James (2001). The Rise of the Public in Enlightenment Europe. Cambridge University Press. pp. 82–83. doi:10.1017/CBO9780511819421. ISBN 978-0511819421. Archived from the original on January 20, 2022. Retrieved May 27, 2022. "The Scientific Revolution and the Enlightenment (1500–1780)" (PDF). Retrieved January 29, 2024. "Scientific Revolution | Definition, History, Scientists, Inventions, & Facts". Britannica. Retrieved January 29, 2024. Madigan M, Martinko J, eds. (2006). Brock Biology of Microorganisms (11th ed.). Prentice Hall. ISBN 978-0131443297. Guicciardini, N. (1999). Reading the Principia: The Debate on Newton's Methods for Natural Philosophy from 1687 to 1736. New York: Cambridge University Press. ISBN 978-0521640664. Calisher, CH (2007). "Taxonomy: what's in a name? Doesn't a rose by any other name smell as sweet?". Croatian Medical Journal. 48 (2): 268–270. PMC 2080517. PMID 17436393. Darrigol, Olivier (2000). Electrodynamics from Ampère to Einstein. New York: Oxford University Press. ISBN 0198505949. Olby, R.C.; Cantor, G.N.; Christie, J.R.R.; Hodge, M.J.S. (1990). Companion to the History of Modern Science. London: Routledge. p. 265. Magnusson, Magnus (November 10, 2003). "Review of James Buchan, Capital of the Mind: how Edinburgh Changed the World". New Statesman. Archived from the original on June 6, 2011. Retrieved April 27, 2014. Swingewood, Alan (1970). "Origins of Sociology: The Case of the Scottish Enlightenment". The British Journal of Sociology. 21 (2): 164–180. doi:10.2307/588406. JSTOR 588406. Fry, Michael (1992). Adam Smith's Legacy: His Place in the Development of Modern Economics. Paul Samuelson, Lawrence Klein, Franco Modigliani, James M. Buchanan, Maurice Allais, Theodore Schultz, Richard Stone, James Tobin, Wassily Leontief, Jan Tinbergen. Routledge. ISBN 978-0-415-06164-3. Lightman, Bernard (2011). "13. Science and the Public". In Shank, Michael; Numbers, Ronald; Harrison, Peter (eds.). Wrestling with Nature: From Omens to Science. Chicago: University of Chicago Press. p. 367. ISBN 978-0-226-31783-0. Leahey, Thomas Hardy (2018). "The psychology of consciousness". A History of Psychology: From Antiquity to Modernity (8th ed.). New York, NY: Routledge. pp. 219–253. ISBN 978-1-138-65242-2. Padian, Kevin (2008). "Darwin's enduring legacy". Nature. 451 (7179): 632–634. Bibcode:2008Natur.451..632P. doi:10.1038/451632a. PMID 18256649. Henig, Robin Marantz (2000). The monk in the garden: the lost and found genius of Gregor Mendel, the father of genetics. pp. 134–138. Miko, Ilona (2008). "Gregor Mendel's principles of inheritance form the cornerstone of modern genetics. So just what are they?". Nature Education. 1 (1): 134. Archived from the original on July 19, 2019. Retrieved May 9, 2021. Rocke, Alan J. (2005). "In Search of El Dorado: John Dalton and the Origins of the Atomic Theory". Social Research. 72 (1): 125–158. doi:10.1353/sor.2005.0003. JSTOR 40972005. S2CID 141350239. Reichl, Linda (1980). A Modern Course in Statistical Physics. Edward Arnold. ISBN 0-7131-2789-9. Rao, Y. V. C. (1997). Chemical Engineering Thermodynamics. Universities Press. p. 158. ISBN 978-81-7371-048-3. Heidrich, M. (2016). "Bounded energy exchange as an alternative to the third law of thermodynamics". Annals of Physics. 373: 665–681. Bibcode:2016AnPhy.373..665H. doi:10.1016/j.aop.2016.07.031. Archived from the original on January 15, 2019. Retrieved May 29, 2022. Mould, Richard F. (1995). A century of X-rays and radioactivity in medicine: with emphasis on photographic records of the early years (Reprint. with minor corr ed.). Bristol: Inst. of Physics Publ. p. 12. ISBN 978-0-7503-0224-1. Estreicher, Tadeusz (1938). "Curie, Maria ze Skłodowskich". Polski słownik biograficzny, vol. 4 (in Polish). p. 113. Thomson, J.J. (1897). "Cathode Rays". Philosophical Magazine. 44 (269): 293–316. doi:10.1080/14786449708621070. Retrieved February 24, 2022. Goyotte, Dolores (2017). "The Surgical Legacy of World War II. Part II: The age of antibiotics" (PDF). The Surgical Technologist. 109: 257–264. Archived (PDF) from the original on May 5, 2021. Retrieved January 8, 2021. Erisman, Jan Willem; MA Sutton; J Galloway; Z Klimont; W Winiwarter (October 2008). "How a century of ammonia synthesis changed the world". Nature Geoscience. 1 (10): 636–639. Bibcode:2008NatGe...1..636E. doi:10.1038/ngeo325. S2CID 94880859. Archived from the original on July 23, 2010. Retrieved October 22, 2010. Emmett, Robert; Zelko, Frank (2014). Emmett, Rob; Zelko, Frank (eds.). "Minding the Gap: Working Across Disciplines in Environmental Studies". Environment & Society Portal. RCC Perspectives no. 2. doi:10.5282/rcc/6313. Archived from the original on January 21, 2022. Furner, Jonathan (June 1, 2003). "Little Book, Big Book: Before and After Little Science, Big Science: A Review Article, Part I". Journal of Librarianship and Information Science. 35 (2): 115–125. doi:10.1177/0961000603352006. S2CID 34844169. Kraft, Chris; James Schefter (2001). Flight: My Life in Mission Control. New York: Dutton. pp. 3–5. ISBN 0-525-94571-7. Kahn, Herman (1962). Thinking about the Unthinkable. Horizon Press. Shrum, Wesley (2007). Structures of scientific collaboration. Joel Genuth, Ivan Chompalov. Cambridge, Mass.: MIT Press. ISBN 978-0-262-28358-8. OCLC 166143348. Archived from the original on July 30, 2022. Retrieved May 31, 2022. Rosser, Sue V. (March 12, 2012). Breaking into the Lab: Engineering Progress for Women in Science. New York: New York University Press. p. 7. ISBN 978-0-8147-7645-2. Penzias, A. A. (2006). "The origin of elements" (PDF). Science. 205 (4406). Nobel Foundation: 549–54. doi:10.1126/science.205.4406.549. PMID 17729659. Archived (PDF) from the original on January 17, 2011. Retrieved October 4, 2006. Weinberg, S. (1972). Gravitation and Cosmology. John Whitney & Sons. pp. 495–464. ISBN 978-0-471-92567-5. Futuyma, Douglas J.; Kirkpatrick, Mark (2017). "Chapter 1: Evolutionary Biology". Evolution (4th ed.). Sinauer. pp. 3–26. ISBN 978-1605356051. Archived from the original on May 31, 2022. Retrieved May 30, 2022. Miller, Arthur I. (1981). Albert Einstein's special theory of relativity. Emergence (1905) and early interpretation (1905–1911). Reading: Addison–Wesley. ISBN 978-0-201-04679-3. ter Haar, D. (1967). The Old Quantum Theory. Pergamon Press. pp. 206. ISBN 978-0-08-012101-7. von Bertalanffy, Ludwig (1972). "The History and Status of General Systems Theory". The Academy of Management Journal. 15 (4): 407–26. JSTOR 255139. Naidoo, Nasheen; Pawitan, Yudi; Soong, Richie; Cooper, David N.; Ku, Chee-Seng (October 2011). "Human genetics and genomics a decade after the release of the draft sequence of the human genome". Human Genomics. 5 (6): 577–622. doi:10.1186/1479-7364-5-6-577. PMC 3525251. PMID 22155605. Rashid, S. Tamir; Alexander, Graeme J.M. (March 2013). "Induced pluripotent stem cells: from Nobel Prizes to clinical applications". Journal of Hepatology. 58 (3): 625–629. doi:10.1016/j.jhep.2012.10.026. ISSN 1600-0641. PMID 23131523. O'Luanaigh, C. (March 14, 2013). "New results indicate that new particle is a Higgs boson" (Press release). CERN. Archived from the original on October 20, 2015. Retrieved October 9, 2013. Abbott, B.P.; Abbott, R.; Abbott, T.D.; Acernese, F.; Ackley, K.; Adams, C.; Adams, T.; Addesso, P.; Adhikari, R.X.; Adya, V.B.; Affeldt, C.; Afrough, M.; Agarwal, B.; Agathos, M.; Agatsuma, K.; Aggarwal, N.; Aguiar, O.D.; Aiello, L.; Ain, A.; Ajith, P.; Allen, B.; Allen, G.; Allocca, A.; Altin, P.A.; Amato, A.; Ananyeva, A.; Anderson, S.B.; Anderson, W.G.; Angelova, S.V.; et al. (2017). "Multi-messenger Observations of a Binary Neutron Star Merger". The Astrophysical Journal. 848 (2): L12. arXiv:1710.05833. Bibcode:2017ApJ...848L..12A. doi:10.3847/2041-8213/aa91c9. S2CID 217162243. Cho, Adrian (2017). "Merging neutron stars generate gravitational waves and a celestial light show". Science. doi:10.1126/science.aar2149. "Media Advisory: First Results from the Event Horizon Telescope to be Presented on April 10th | Event Horizon Telescope". April 20, 2019. Archived from the original on April 20, 2019. Retrieved September 21, 2021. "Scientific Method: Relationships Among Scientific Paradigms". Seed Magazine. March 7, 2007. Archived from the original on November 1, 2016. Retrieved November 4, 2016. Bunge, Mario Augusto (1998). Philosophy of Science: From Problem to Theory. Transaction Publishers. p. 24. ISBN 978-0-7658-0413-6. Popper, Karl R. (2002a) [1959]. "A survey of some fundamental problems". The Logic of Scientific Discovery. New York: Routledge Classics. pp. 3–26. ISBN 978-0-415-27844-7. OCLC 59377149. Gauch, Hugh G. Jr. (2003). "Science in perspective". Scientific Method in Practice. Cambridge, United Kingdom: Cambridge University Press. pp. 21–73. ISBN 978-0-521-01708-4. Archived from the original on December 25, 2020. Retrieved September 3, 2018. Oglivie, Brian W. (2008). "Introduction". The Science of Describing: Natural History in Renaissance Europe (Paperback ed.). Chicago: University of Chicago Press. pp. 1–24. ISBN 978-0-226-62088-6. "Natural History". Princeton University WordNet. Archived from the original on March 3, 2012. Retrieved October 21, 2012. "Formal Sciences: Washington and Lee University". Washington and Lee University. Archived from the original on May 14, 2021. Retrieved May 14, 2021. A "formal science" is an area of study that uses formal systems to generate knowledge such as in Mathematics and Computer Science. Formal sciences are important subjects because all of quantitative science depends on them. "formal system". Encyclopædia Britannica. Archived from the original on April 29, 2008. Retrieved May 30, 2022. Tomalin, Marcus (2006). Linguistics and the Formal Sciences. Löwe, Benedikt (2002). "The Formal Sciences: Their Scope, Their Foundations, and Their Unity". Synthese. 133: 5–11. doi:10.1023/a:1020887832028. S2CID 9272212. Bill, Thompson (2007). "2.4 Formal Science and Applied Mathematics". The Nature of Statistical Evidence. Lecture Notes in Statistics. Vol. 189. Springer. p. 15. Bunge, Mario (1998). "The Scientific Approach". Philosophy of Science: Volume 1, From Problem to Theory. Vol. 1 (revised ed.). New York: Routledge. pp. 3–50. ISBN 978-0-7658-0413-6. Mujumdar, Anshu Gupta; Singh, Tejinder (2016). "Cognitive science and the connection between physics and mathematics". In Aguirre, Anthony; Foster, Brendan (eds.). Trick or Truth?: The Mysterious Connection Between Physics and Mathematics. The Frontiers Collection. Switzerland: SpringerNature. pp. 201–218. ISBN 978-3-319-27494-2. "About the Journal". Journal of Mathematical Physics. Archived from the original on October 3, 2006. Retrieved October 3, 2006. Restrepo, G. (2016). "Mathematical chemistry, a new discipline". In Scerri, E.; Fisher, G. (eds.). Essays in the philosophy of chemistry. New York, UK: Oxford University Press. pp. 332–351. ISBN 978-0-19-049459-9. Archived from the original on June 10, 2021. "What is mathematical biology". Centre for Mathematical Biology, University of Bath. Archived from the original on September 23, 2018. Retrieved June 7, 2018. Johnson, Tim (September 1, 2009). "What is financial mathematics?". +Plus Magazine. Archived from the original on April 8, 2022. Retrieved March 1, 2021. Varian, Hal (1997). "What Use Is Economic Theory?". In D'Autume, A.; Cartelier, J. (eds.). Is Economics Becoming a Hard Science?. Edward Elgar. Pre-publication. Archived June 25, 2006, at the Wayback Machine. Retrieved April 1, 2008. Abraham, Reem Rachel (2004). "Clinically oriented physiology teaching: strategy for developing critical-thinking skills in undergraduate medical students". Advances in Physiology Education. 28 (3): 102–04. doi:10.1152/advan.00001.2004. PMID 15319191. S2CID 21610124. "Cambridge Dictionary". Cambridge University Press. Archived from the original on August 19, 2019. Retrieved March 25, 2021. Brooks, Harvey (September 1, 1994). "The relationship between science and technology" (PDF). Research Policy. Special Issue in Honor of Nathan Rosenberg. 23 (5): 477–486. doi:10.1016/0048-7333(94)01001-3. ISSN 0048-7333. Firth, John (2020). "Science in medicine: when, how, and what". Oxford textbook of medicine. Oxford: Oxford University Press. ISBN 978-0-19-874669-0. Saunders, J. (June 2000). "The practice of clinical medicine as an art and as a science". Med Humanit. 26 (1): 18–22. doi:10.1136/mh.26.1.18. PMC 1071282. PMID 12484313. S2CID 73306806. Davis, Bernard D. (March 2000). "Limited scope of science". Microbiology and Molecular Biology Reviews. 64 (1): 1–12. doi:10.1128/MMBR.64.1.1-12.2000. PMC 98983. PMID 10704471 & "Technology" in Davis, Bernard (March 2000). "The scientist's world". Microbiology and Molecular Biology Reviews. 64 (1): 1–12. doi:10.1128/MMBR.64.1.1-12.2000. PMC 98983. PMID 10704471. McCormick, James (2001). "Scientific medicine—fact of fiction? The contribution of science to medicine". Occasional Paper (Royal College of General Practitioners) (80): 3–6. PMC 2560978. PMID 19790950. Breznau, Nate (2022). "Integrating Computer Prediction Methods in Social Science: A Comment on Hofman et al. (2021)". Social Science Computer Review. 40 (3): 844–853. doi:10.1177/08944393211049776. S2CID 248334446. Hofman, Jake M.; Watts, Duncan J.; Athey, Susan; Garip, Filiz; Griffiths, Thomas L.; Kleinberg, Jon; Margetts, Helen; Mullainathan, Sendhil; Salganik, Matthew J.; Vazire, Simine; Vespignani, Alessandro (July 2021). "Integrating explanation and prediction in computational social science". Nature. 595 (7866): 181–188. Bibcode:2021Natur.595..181H. doi:10.1038/s41586-021-03659-0. ISSN 1476-4687. PMID 34194044. S2CID 235697917. Archived from the original on September 25, 2021. Retrieved September 25, 2021. Nissani, M. (1995). "Fruits, Salads, and Smoothies: A Working definition of Interdisciplinarity". The Journal of Educational Thought. 29 (2): 121–128. JSTOR 23767672. Moody G (2004). Digital Code of Life: How Bioinformatics is Revolutionizing Science, Medicine, and Business. John Wiley & Sons. p. vii. ISBN 978-0-471-32788-2. Ausburg, Tanya (2006). Becoming Interdisciplinary: An Introduction to Interdisciplinary Studies (2nd ed.). New York: Kendall/Hunt Publishing. Dawkins, Richard (May 10, 2006). "To Live at All Is Miracle Enough". RichardDawkins.net. Archived from the original on January 19, 2012. Retrieved February 5, 2012. di Francia, Giuliano Toraldo (1976). "The method of physics". The Investigation of the Physical World. Cambridge, United Kingdom: Cambridge University Press. pp. 1–52. ISBN 978-0-521-29925-1. The amazing point is that for the first time since the discovery of mathematics, a method has been introduced, the results of which have an intersubjective value! Popper, Karl R. (2002e) [1959]. "The problem of the empirical basis". The Logic of Scientific Discovery. New York: Routledge Classics. pp. 3–26. ISBN 978-0-415-27844-7. OCLC 59377149. Diggle, Peter J.; Chetwynd, Amanda G. (2011). Statistics and Scientific Method: An Introduction for Students and Researchers. Oxford University Press. pp. 1, 2. ISBN 978-0199543182. Wilson, Edward (1999). Consilience: The Unity of Knowledge. New York: Vintage. ISBN 978-0-679-76867-8. Fara, Patricia (2009). "Decisions". Science: A Four Thousand Year History. Oxford, United Kingdom: Oxford University Press. p. 408. ISBN 978-0-19-922689-4. Aldrich, John (1995). "Correlations Genuine and Spurious in Pearson and Yule". Statistical Science. 10 (4): 364–376. doi:10.1214/ss/1177009870. JSTOR 2246135. Nola, Robert; Irzik, Gürol (2005k). "naive inductivism as a methodology in science". Philosophy, science, education and culture. Science & technology education library. Vol. 28. Springer. pp. 207–230. ISBN 978-1-4020-3769-6. Nola, Robert; Irzik, Gürol (2005j). "The aims of science and critical inquiry". Philosophy, science, education and culture. Science & technology education library. Vol. 28. Springer. pp. 207–230. ISBN 978-1-4020-3769-6. van Gelder, Tim (1999). ""Heads I win, tails you lose": A Foray Into the Psychology of Philosophy" (PDF). University of Melbourne. Archived from the original (PDF) on April 9, 2008. Retrieved March 28, 2008. Pease, Craig (September 6, 2006). "Chapter 23. Deliberate bias: Conflict creates bad science". Science for Business, Law and Journalism. Vermont Law School. Archived from the original on June 19, 2010. Shatz, David (2004). Peer Review: A Critical Inquiry. Rowman & Littlefield. ISBN 978-0-7425-1434-8. OCLC 54989960. Krimsky, Sheldon (2003). Science in the Private Interest: Has the Lure of Profits Corrupted the Virtue of Biomedical Research. Rowman & Littlefield. ISBN 978-0-7425-1479-9. OCLC 185926306. Bulger, Ruth Ellen; Heitman, Elizabeth; Reiser, Stanley Joel (2002). The Ethical Dimensions of the Biological and Health Sciences (2nd ed.). Cambridge University Press. ISBN 978-0-521-00886-0. OCLC 47791316. Backer, Patricia Ryaby (October 29, 2004). "What is the scientific method?". San Jose State University. Archived from the original on April 8, 2008. Retrieved March 28, 2008. Ziman, John (1978c). "Common observation". Reliable knowledge: An exploration of the grounds for belief in science. Cambridge: Cambridge University Press. pp. 42–76. ISBN 978-0-521-22087-3. Ziman, J.M. (1980). "The proliferation of scientific literature: a natural process". Science. 208 (4442): 369–71. Bibcode:1980Sci...208..369Z. doi:10.1126/science.7367863. PMID 7367863. Subramanyam, Krishna; Subramanyam, Bhadriraju (1981). Scientific and Technical Information Resources. CRC Press. ISBN 978-0-8247-8297-9. OCLC 232950234. Bush, Vannevar (July 1945). "Science the Endless Frontier". National Science Foundation. Archived from the original on November 7, 2016. Retrieved November 4, 2016. Schooler, J. W. (2014). "Metascience could rescue the 'replication crisis'". Nature. 515 (7525): 9. Bibcode:2014Natur.515....9S. doi:10.1038/515009a. PMID 25373639. Pashler, Harold; Wagenmakers, Eric Jan (2012). "Editors' Introduction to the Special Section on Replicability in Psychological Science: A Crisis of Confidence?". Perspectives on Psychological Science. 7 (6): 528–530. doi:10.1177/1745691612465253. PMID 26168108. S2CID 26361121. Ioannidis, John P. A.; Fanelli, Daniele; Dunne, Debbie Drake; Goodman, Steven N. (October 2, 2015). "Meta-research: Evaluation and Improvement of Research Methods and Practices". PLOS Biology. 13 (10): –1002264. doi:10.1371/journal.pbio.1002264. ISSN 1545-7885. PMC 4592065. PMID 26431313. Hansson, Sven Ove; Zalta, Edward N. (September 3, 2008). "Science and Pseudoscience". Stanford Encyclopedia of Philosophy. Section 2: The "science" of pseudoscience. Archived from the original on October 29, 2021. Retrieved May 28, 2022. Shermer M (1997). Why people believe weird things: pseudoscience, superstition, and other confusions of our time. New York: W. H. Freeman and Company. p. 17. ISBN 978-0-7167-3090-3. Feynman, Richard (1974). "Cargo Cult Science". Center for Theoretical Neuroscience. Columbia University. Archived from the original on March 4, 2005. Retrieved November 4, 2016. Novella, Steven (2018). The Skeptics' Guide to the Universe: How to Know What's Really Real in a World Increasingly Full of Fake. Hodder & Stoughton. p. 162. ISBN 978-1473696419. "Coping with fraud" (PDF). The COPE Report 1999: 11–18. Archived from the original (PDF) on September 28, 2007. Retrieved July 21, 2011. It is 10 years, to the month, since Stephen Lock ... Reproduced with kind permission of the Editor, The Lancet. Godfrey-Smith, Peter (2003c). "Induction and confirmation". Theory and Reality: An Introduction to the Philosophy of Science. Chicago: University of Chicago. pp. 39–56. ISBN 978-0-226-30062-7. Godfrey-Smith, Peter (2003o). "Empiricism, naturalism, and scientific realism?". Theory and Reality: An Introduction to the Philosophy of Science. Chicago: University of Chicago. pp. 219–232. ISBN 978-0-226-30062-7. Godfrey-Smith, Peter (2003b). "Logic plus empiricism". Theory and Reality: An Introduction to the Philosophy of Science. Chicago: University of Chicago. pp. 19–38. ISBN 978-0-226-30062-7. Godfrey-Smith, Peter (2003d). "Popper: Conjecture and refutation". Theory and Reality: An Introduction to the Philosophy of Science. Chicago: University of Chicago. pp. 57–74. ISBN 978-0-226-30062-7. Godfrey-Smith, Peter (2003g). "Lakatos, Laudan, Feyerabend, and frameworks". Theory and Reality: An Introduction to the Philosophy of Science. Chicago: University of Chicago. pp. 102–121. ISBN 978-0-226-30062-7. Popper, Karl (1972). Objective Knowledge. Newton-Smith, W.H. (1994). The Rationality of Science. London: Routledge. p. 30. ISBN 978-0-7100-0913-5. Votsis, I. (2004). The Epistemological Status of Scientific Theories: An Investigation of the Structural Realist Account (PhD Thesis). University of London, London School of Economics. p. 39. Bird, Alexander (2013). Zalta, Edward N. (ed.). "Thomas Kuhn". Stanford Encyclopedia of Philosophy. Archived from the original on July 15, 2020. Retrieved October 26, 2015. Kuhn, Thomas S. (1970). The Structure of Scientific Revolutions (2nd ed.). University of Chicago Press. p. 206. ISBN 978-0-226-45804-5. Archived from the original on October 19, 2021. Retrieved May 30, 2022. Godfrey-Smith, Peter (2003). "Naturalistic philosophy in theory and practice". Theory and Reality: An Introduction to the Philosophy of Science. Chicago: University of Chicago. pp. 149–162. ISBN 978-0-226-30062-7. Brugger, E. Christian (2004). "Casebeer, William D. Natural Ethical Facts: Evolution, Connectionism, and Moral Cognition". The Review of Metaphysics. 58 (2). Kornfeld, W; Hewitt, CE (1981). "The Scientific Community Metaphor" (PDF). IEEE Transactions on Systems, Man, and Cybernetics. 11 (1): 24–33. doi:10.1109/TSMC.1981.4308575. hdl:1721.1/5693. S2CID 1322857. Archived (PDF) from the original on April 8, 2016. Retrieved May 26, 2022. "Eusocial climbers" (PDF). E.O. Wilson Foundation. Archived (PDF) from the original on April 27, 2019. Retrieved September 3, 2018. But he's not a scientist, he's never done scientific research. My definition of a scientist is that you can complete the following sentence: 'he or she has shown that...'," Wilson says. "Our definition of a scientist". Science Council. Archived from the original on August 23, 2019. Retrieved September 7, 2018. A scientist is someone who systematically gathers and uses research and evidence, making a hypothesis and testing it, to gain and share understanding and knowledge. Cyranoski, David; Gilbert, Natasha; Ledford, Heidi; Nayar, Anjali; Yahia, Mohammed (2011). "Education: The PhD factory". Nature. 472 (7343): 276–79. Bibcode:2011Natur.472..276C. doi:10.1038/472276a. PMID 21512548. Kwok, Roberta (2017). "Flexible working: Science in the gig economy". Nature. 550: 419–21. doi:10.1038/nj7677-549a. Woolston, Chris (2007). Editorial (ed.). "Many junior scientists need to take a hard look at their job prospects". Nature. 550: 549–552. doi:10.1038/nj7677-549a. Lee, Adrian; Dennis, Carina; Campbell, Phillip (2007). "Graduate survey: A love–hurt relationship". Nature. 550 (7677): 549–52. doi:10.1038/nj7677-549a. Cyranoski, David; Gilbert, Natasha; Ledford, Heidi; Nayar, Anjali; Yahia, Mohammed (2011). "Education: The PhD factory". Nature. 472 (7343): 276–279. Bibcode:2011Natur.472..276C. doi:10.1038/472276a. PMID 21512548. Kwok, Roberta (2017). "Flexible working: Science in the gig economy". Nature. 550: 419–421. doi:10.1038/nj7677-549a. Lee, Adrian; Dennis, Carina; Campbell, Phillip (2007). "Graduate survey: A love–hurt relationship". Nature. 550 (7677): 549–552. doi:10.1038/nj7677-549a. Whaley, Leigh Ann (2003). Women's History as Scientists. Santa Barbara, California: ABC-CLIO, INC. Spanier, Bonnie (1995). "From Molecules to Brains, Normal Science Supports Sexist Beliefs about Difference". Im/partial Science: Gender Identity in Molecular Biology. Indiana University Press. ISBN 978-0-253-20968-9. Parrott, Jim (August 9, 2007). "Chronicle for Societies Founded from 1323 to 1599". Scholarly Societies Project. Archived from the original on January 6, 2014. Retrieved September 11, 2007. "The Environmental Studies Association of Canada – What is a Learned Society?". Archived from the original on May 29, 2013. Retrieved May 10, 2013. "Learned societies & academies". Archived from the original on June 3, 2014. Retrieved May 10, 2013. "Learned Societies, the key to realising an open access future?". Impact of Social Sciences. London School of Economics. June 24, 2019. Retrieved January 22, 2023. "Accademia Nazionale dei Lincei" (in Italian). 2006. Archived from the original on February 28, 2010. Retrieved September 11, 2007. "Prince of Wales opens Royal Society's refurbished building". The Royal Society. July 7, 2004. Archived from the original on April 9, 2015. Retrieved December 7, 2009. Meynell, G.G. "The French Academy of Sciences, 1666–91: A reassessment of the French Académie royale des sciences under Colbert (1666–83) and Louvois (1683–91)". Archived from the original on January 18, 2012. Retrieved October 13, 2011. ITS. "Founding of the National Academy of Sciences". .nationalacademies.org. Archived from the original on February 3, 2013. Retrieved March 12, 2012. "The founding of the Kaiser Wilhelm Society (1911)". Max-Planck-Gesellschaft. Archived from the original on March 2, 2022. Retrieved May 30, 2022. "Introduction". Chinese Academy of Sciences. Archived from the original on March 31, 2022. Retrieved May 31, 2022. "Two main Science Councils merge to address complex global challenges". UNESCO. July 5, 2018. Archived from the original on July 12, 2021. Retrieved October 21, 2018. Stockton, Nick (October 7, 2014). "How did the Nobel Prize become the biggest award on Earth?". Wired. Archived from the original on June 19, 2019. Retrieved September 3, 2018. "Main Science and Technology Indicators – 2008-1" (PDF). OECD. Archived from the original (PDF) on February 15, 2010. OECD Science, Technology and Industry Scoreboard 2015: Innovation for growth and society. OECD. 2015. p. 156. doi:10.1787/sti_scoreboard-2015-en. ISBN 978-9264239784. Archived from the original on May 25, 2022. Retrieved May 28, 2022 – via oecd-ilibrary.org. Kevles, Daniel (1977). "The National Science Foundation and the Debate over Postwar Research Policy, 1942-1945". Isis. 68 (241): 4–26. doi:10.1086/351711. PMID 320157. S2CID 32956693. "Argentina, National Scientific and Technological Research Council (CONICET)". International Science Council. Archived from the original on May 16, 2022. Retrieved May 31, 2022. Innis, Michelle (May 17, 2016). "Australia to Lay Off Leading Scientist on Sea Levels". The New York Times. ISSN 0362-4331. Archived from the original on May 7, 2021. Retrieved May 31, 2022. "Le CNRS recherche 10.000 passionnés du blob". Le Figaro (in French). October 20, 2021. Archived from the original on April 27, 2022. Retrieved May 31, 2022. Bredow, Rafaela von (December 18, 2021). "How a Prestigious Scientific Organization Came Under Suspicion of Treating Women Unequally". Der Spiegel. ISSN 2195-1349. Archived from the original on May 29, 2022. Retrieved May 31, 2022. "En espera de una "revolucionaria" noticia sobre Sagitario A*, el agujero negro supermasivo en el corazón de nuestra galaxia". ELMUNDO (in Spanish). May 12, 2022. Archived from the original on May 13, 2022. Retrieved May 31, 2022. Fletcher, Anthony C.; Bourne, Philip E. (September 27, 2012). "Ten Simple Rules To Commercialize Scientific Research". PLOS Computational Biology. 8 (9): e1002712. Bibcode:2012PLSCB...8E2712F. doi:10.1371/journal.pcbi.1002712. ISSN 1553-734X. PMC 3459878. PMID 23028299. Marburger, John Harmen III (February 10, 2015). Science policy up close. Crease, Robert P. Cambridge, Massachusetts: Harvard University Press. ISBN 978-0-674-41709-0. OCLC 875999943. Benneworth, Paul; Jongbloed, Ben W. (July 31, 2009). "Who matters to universities? A stakeholder perspective on humanities, arts and social sciences valorisation" (PDF). Higher Education. 59 (5): 567–588. doi:10.1007/s10734-009-9265-2. ISSN 0018-1560. Dickson, David (October 11, 2004). "Science journalism must keep a critical edge". Science and Development Network. Archived from the original on June 21, 2010. Mooney, Chris (November–December 2004). "Blinded By Science, How 'Balanced' Coverage Lets the Scientific Fringe Hijack Reality". Columbia Journalism Review. Vol. 43, no. 4. Archived from the original on January 17, 2010. Retrieved February 20, 2008. McIlwaine, S.; Nguyen, D.A. (2005). "Are Journalism Students Equipped to Write About Science?". Australian Studies in Journalism. 14: 41–60. Archived from the original on August 1, 2008. Retrieved February 20, 2008. Webb, Sarah (December 2013). "Popular science: Get the word out". Nature. 504 (7478): 177–9. doi:10.1038/nj7478-177a. PMID 24312943. Wilde, Fran (January 21, 2016). "How Do You Like Your Science Fiction? Ten Authors Weigh In On 'Hard' vs. 'Soft' SF". Tor.com. Archived from the original on April 4, 2019. Retrieved April 4, 2019. Petrucci, Mario. "Creative Writing – Science". Archived from the original on January 6, 2009. Retrieved April 27, 2008. Tyson, Alec; Funk, Cary; Kennedy, Brian; Johnson, Courtney (September 15, 2021). "Majority in U.S. Says Public Health Benefits of COVID-19 Restrictions Worth the Costs, Even as Large Shares Also See Downsides". Pew Research Center Science & Society. Retrieved August 4, 2022. Kennedy, Brian. "U.S. concern about climate change is rising, but mainly among Democrats". Pew Research Center. Retrieved August 4, 2022. Philipp-Muller, Aviva; Lee, Spike W. S.; Petty, Richard E. (July 26, 2022). "Why are people antiscience, and what can we do about it?". Proceedings of the National Academy of Sciences. 119 (30): e2120755119. Bibcode:2022PNAS..11920755P. doi:10.1073/pnas.2120755119. ISSN 0027-8424. PMC 9335320. PMID 35858405. Gauchat, Gordon William (2008). "A Test of Three Theories of Anti-Science Attitudes". Sociological Focus. 41 (4): 337–357. doi:10.1080/00380237.2008.10571338. S2CID 144645723. Poushter, Jacob; Fagan, Moira; Gubbala, Sneha (August 31, 2022). "Climate Change Remains Top Global Threat Across 19-Country Survey". Pew Research Center's Global Attitudes Project. Retrieved September 5, 2022. McRaney, David (2022). How Minds Change: The Surprising Science of Belief, Opinion, and Persuasion. [New York, NY]: Portfolio/Penguin. ISBN 978-0-593-19029-6. OCLC 1322437138. McGreal, Chris (October 26, 2021). "Revealed: 60% of Americans say oil firms are to blame for the climate crisis". The Guardian. Archived from the original on October 26, 2021. Source: Guardian/Vice/CCN/YouGov poll. Note: ±4% margin of error. Goldberg, Jeanne (2017). "The Politicization of Scientific Issues: Looking through Galileo's Lens or through the Imaginary Looking Glass". Skeptical Inquirer. 41 (5): 34–39. Archived from the original on August 16, 2018. Retrieved August 16, 2018. Bolsen, Toby; Druckman, James N. (2015). "Counteracting the Politicization of Science". Journal of Communication (65): 746. Freudenberg, William F.; Gramling, Robert; Davidson, Debra J. (2008). "Scientific Certainty Argumentation Methods (SCAMs): Science and the Politics of Doubt" (PDF). Sociological Inquiry. 78: 2–38. doi:10.1111/j.1475-682X.2008.00219.x. Archived (PDF) from the original on November 26, 2020. Retrieved April 12, 2020. van der Linden, Sander; Leiserowitz, Anthony; Rosenthal, Seth; Maibach, Edward (2017). "Inoculating the Public against Misinformation about Climate Change" (PDF). Global Challenges. 1 (2): 1. Bibcode:2017GloCh...100008V. doi:10.1002/gch2.201600008. PMC 6607159. PMID 31565263. Archived (PDF) from the original on April 4, 2020. Retrieved August 25, 2019. External links Look up science in Wiktionary, the free dictionary. vte Philosophy of science Concepts AnalysisAnalytic–synthetic distinctionA priori and a posterioriCausalityCommensurabilityConsilienceConstructCreative synthesisDemarcation problemEmpirical evidenceExplanatory powerFactFalsifiabilityFeminist methodFunctional contextualismIgnoramus et ignorabimusInductive reasoningIntertheoretic reductionInquiryNatureObjectivityObservationParadigmProblem of inductionScientific evidence Evidence-based practiceScientific lawScientific methodScientific pluralismScientific RevolutionScientific theoryTestabilityTheory choiceTheory-ladennessUnderdeterminationUnity of sciencemore... Theories CoherentismConfirmation holismConstructive empiricismConstructive realismConstructivist epistemologyContextualismConventionalismDeductive-nomological modelEpistemological anarchismEvolutionismFallibilismFoundationalismHypothetico-deductive modelInductionismInstrumentalismModel-dependent realismNaturalismPhysicalismPositivism / Reductionism / DeterminismPragmatismRationalism / EmpiricismReceived view / Semantic view of theoriesScientific essentialismScientific formalismScientific realism / Anti-realismScientific skepticismScientismStructuralismUniformitarianismVitalism Philosophy of... BiologyChemistryPhysics Space and timeSocial science ArchaeologyEconomics‎GeographyHistoryLinguisticsPsychology Related topics Criticism of scienceDescriptive scienceEpistemologyFaith and rationalityHard and soft scienceHistory and philosophy of scienceNormative scienceProtosciencePseudoscienceRelationship between religion and scienceRhetoric of scienceScience studiesSociology of scientific ignoranceSociology of scientific knowledge Philosophers of science Precursors Roger BaconFrancis BaconGalileo GalileiIsaac Newton Auguste ComteHenri PoincaréPierre DuhemRudolf SteinerKarl PearsonCharles Sanders PeirceWilhelm WindelbandAlfred North WhiteheadBertrand RussellOtto NeurathC. D. BroadMichael PolanyiHans ReichenbachRudolf CarnapKarl PopperCarl Gustav HempelW. V. O. QuineThomas KuhnImre LakatosPaul FeyerabendIan HackingBas van FraassenLarry Laudan Category Philosophy portalicon Science portal vte Glossaries of science and engineering Aerospace engineeringAgricultureArchaeologyArchitectureArtificial intelligenceAstronomyBiologyBotanyCalculusCell biologyChemistryCivil engineeringClinical researchComputer hardwareComputer scienceDevelopmental and reproductive biologyEcologyEconomicsElectrical and electronics engineeringEngineering A–LM–ZEntomologyEnvironmental scienceGenetics and evolutionary biologyCellular and molecular biology 0–LM–ZGeography A–MN–ZArabic toponymsHebrew toponymsWestern and South AsiaGeologyIchthyologyMachine visionMathematicsMechanical engineeringMedicineMeteorologyMycologyNanotechnologyOrnithologyPhysicsProbability and statisticsPsychiatryQuantum computingRoboticsScientific namingStructural engineeringVirology vte Science and technology studies Economics Economics of scienceEconomics of scientific knowledge History History and philosophy of scienceHistory of science and technologyHistory of technology Philosophy AnthropoceneAntipositivismEmpiricismFuzzy logicNeo-LuddismPhilosophy of sciencePhilosophy of social sciencePhilosophy of technologyPositivismPostpositivismReligion and scienceScientismSocial constructivismSocial epistemologyTranshumanism Sociology Actor–network theorySocial construction of technologyshaping of technologySociology of knowledge scientificSociology of scientific ignoranceSociology of the history of scienceSociotechnologyStrong programme Science studies AntiscienceBibliometricsBoundary-workConsilienceCriticism of scienceDemarcation problemDouble hermeneuticLogologyMapping controversiesMetascienceParadigm shift black swan eventsPseudosciencePsychology of scienceScience citizencommunicationeducationnormalNeo-colonialpost-normalrhetoricwarsScientific communityconsensuscontroversydissententerpriseliteracymethodmisconductpriorityskepticismScientocracyScientometricsTeam scienceTraditional knowledge ecologicalUnity of scienceWomen in science STEM Technology studies Co-productionCyborg anthropologyDesign studiesDematerializationDigital anthropologyDigital media use and mental healthEarly adopterEngineering studiesFinancial technologyHype cycleInnovation diffusiondisruptivelinear modelsystemuserLeapfroggingNormalization process theoryMedia studiesReverse salientSkunkworks projectSociotechnical systemTechnical changeTechnocracyTechnoscience feministTechnological changeconvergencedeterminismrevolutiontransitionsTechnology and societycriticism ofdynamicstheories oftransferWomen in engineering Policy Academic freedomDigital divideEvidence-based policyFactor 10Funding of scienceHorizon scanningPoliticization of scienceRegulation of scienceResearch ethicsRight to scienceScience policy history ofscience ofTechnology assessmentTechnology policyTransition management Portals ScienceHistory of scienceTechnology Category AssociationsJournalsScholars Portal: icon Science Science at Wikipedia's sister projects: Media from Commons News from Wikinews Quotations from Wikiquote Textbooks from Wikibooks Resources from Wikiversity Travel guides from Wikivoyage Categories: ScienceObservationMain topic articles Science Article Talk Read View source View history Tools Page semi-protected From Wikipedia, the free encyclopedia For other uses, see Science (disambiguation). For a topical guide, see Outline of science. Part of a series on Science icon Science portalOutlineCategoryIndexGlossaryDisambiguationHistoryLiteraturePhilosophy Fields (Outline / List) Intrascientific fields Applied sciencesFormal sciences MathematicalComputerInterdisciplinary sciencesNatural sciences PhysicalLifeEnvironmentalSocial sciences CulturalEconomicalHumanPolitical Extrascientific fields ArtsCommunication studiesCraftFuturologyHistoryHumanitiesKnowledge managementLanguage studiesLawLiberal artsLiteratureMusicPhilosophyPolemologyProfessionsReligionResearch and developmentStrategic studiesUrban studiesVocational education Scientific integrity ReproducibilityCognitive biasLogical fallacyResearch ethics Instruments Science communicationScience educationResearch fundingScientific methodScience policyScientistResearchTechnology This is a subseries on philosophy. In order to explore related topics, please visit navigation. vte Science is a rigorous, systematic endeavor that builds and organizes knowledge in the form of testable explanations and predictions about the world.[1][2] Modern science is typically divided into three major branches:[3] the natural sciences (e.g., physics, chemistry, and biology), which study the physical world; the social sciences (e.g., economics, psychology, and sociology), which study individuals and societies;[4][5] and the formal sciences (e.g., logic, mathematics, and theoretical computer science), which study formal systems, governed by axioms and rules.[6][7] There is disagreement whether the formal sciences are science disciplines,[8][9][10] as they do not rely on empirical evidence.[11][9] Applied sciences are disciplines that use scientific knowledge for practical purposes, such as in engineering and medicine.[12][13][14] The history of science spans the majority of the historical record, with the earliest written records of identifiable predecessors to modern science dating to Bronze Age Egypt and Mesopotamia from around 3000 to 1200 BCE. Their contributions to mathematics, astronomy, and medicine entered and shaped the Greek natural philosophy of classical antiquity, whereby formal attempts were made to provide explanations of events in the physical world based on natural causes, while further advancements, including the introduction of the Hindu–Arabic numeral system, were made during the Golden Age of India.[15]: 12 [16][17][18] Scientific research deteriorated in these regions after the fall of the Western Roman Empire during the Early Middle Ages (400 to 1000 CE), but in the Medieval renaissances (Carolingian Renaissance, Ottonian Renaissance and the Renaissance of the 12th century) scholarship flourished again. Some Greek manuscripts lost in Western Europe were preserved and expanded upon in the Middle East during the Islamic Golden Age,[19] along with the later efforts of Byzantine Greek scholars who brought Greek manuscripts from the dying Byzantine Empire to Western Europe at the start of the Renaissance. The recovery and assimilation of Greek works and Islamic inquiries into Western Europe from the 10th to 13th century revived "natural philosophy",[20][21][22] which was later transformed by the Scientific Revolution that began in the 16th century[23] as new ideas and discoveries departed from previous Greek conceptions and traditions.[24][25] The scientific method soon played a greater role in knowledge creation and it was not until the 19th century that many of the institutional and professional features of science began to take shape,[26][27] along with the changing of "natural philosophy" to "natural science".[28] New knowledge in science is advanced by research from scientists who are motivated by curiosity about the world and a desire to solve problems.[29][30] Contemporary scientific research is highly collaborative and is usually done by teams in academic and research institutions,[31] government agencies, and companies.[32][33] The practical impact of their work has led to the emergence of science policies that seek to influence the scientific enterprise by prioritizing the ethical and moral development of commercial products, armaments, health care, public infrastructure, and environmental protection. Etymology The word science has been used in Middle English since the 14th century in the sense of "the state of knowing". The word was borrowed from the Anglo-Norman language as the suffix -cience, which was borrowed from the Latin word scientia, meaning "knowledge, awareness, understanding". It is a noun derivative of the Latin sciens meaning "knowing", and undisputedly derived from the Latin sciō, the present participle scīre, meaning "to know".[34] There are many hypotheses for science's ultimate word origin. According to Michiel de Vaan, Dutch linguist and Indo-Europeanist, sciō may have its origin in the Proto-Italic language as *skije- or *skijo- meaning "to know", which may originate from Proto-Indo-European language as *skh1-ie, *skh1-io, meaning "to incise". The Lexikon der indogermanischen Verben proposed sciō is a back-formation of nescīre, meaning "to not know, be unfamiliar with", which may derive from Proto-Indo-European *sekH- in Latin secāre, or *skh2-, from *sḱʰeh2(i)- meaning "to cut".[35] In the past, science was a synonym for "knowledge" or "study", in keeping with its Latin origin. A person who conducted scientific research was called a "natural philosopher" or "man of science".[36] In 1834, William Whewell introduced the term scientist in a review of Mary Somerville's book On the Connexion of the Physical Sciences,[37] crediting it to "some ingenious gentleman" (possibly himself).[38] History Main article: History of science Early history Main article: History of science in early cultures Clay tablet with markings, three columns for numbers and one for ordinals The Plimpton 322 tablet by the Babylonians records Pythagorean triples, written in about 1800 BCE Science has no single origin. Rather, systematic methods emerged gradually over the course of tens of thousands of years,[39][40] taking different forms around the world, and few details are known about the very earliest developments. Women likely played a central role in prehistoric science,[41] as did religious rituals.[42] Some scholars use the term "protoscience" to label activities in the past that resemble modern science in some but not all features;[43][44][45] however, this label has also been criticized as denigrating,[46] or too suggestive of presentism, thinking about those activities only in relation to modern categories.[47] Direct evidence for scientific processes becomes clearer with the advent of writing systems in early civilizations like Ancient Egypt and Mesopotamia, creating the earliest written records in the history of science in around 3000 to 1200 BCE.[15]: 12–15 [16] Although the words and concepts of "science" and "nature" were not part of the conceptual landscape at the time, the ancient Egyptians and Mesopotamians made contributions that would later find a place in Greek and medieval science: mathematics, astronomy, and medicine.[48][15]: 12  From the 3rd millennium BCE, the ancient Egyptians developed a decimal numbering system,[49] solved practical problems using geometry,[50] and developed a calendar.[51] Their healing therapies involved drug treatments and the supernatural, such as prayers, incantations, and rituals.[15]: 9  The ancient Mesopotamians used knowledge about the properties of various natural chemicals for manufacturing pottery, faience, glass, soap, metals, lime plaster, and waterproofing.[52] They studied animal physiology, anatomy, behavior, and astrology for divinatory purposes.[53] The Mesopotamians had an intense interest in medicine and the earliest medical prescriptions appeared in Sumerian during the Third Dynasty of Ur.[52][54] They seem to have studied scientific subjects which had practical or religious applications and had little interest in satisfying curiosity.[52] Classical antiquity Main article: Science in classical antiquity Framed mosaic of philosophers gathering around and conversing Plato's Academy mosaic, made between 100 BCE to 79 AD, shows many Greek philosophers and scholars In classical antiquity, there is no real ancient analog of a modern scientist. Instead, well-educated, usually upper-class, and almost universally male individuals performed various investigations into nature whenever they could afford the time.[55] Before the invention or discovery of the concept of phusis or nature by the pre-Socratic philosophers, the same words tend to be used to describe the natural "way" in which a plant grows,[56] and the "way" in which, for example, one tribe worships a particular god. For this reason, it is claimed that these men were the first philosophers in the strict sense and the first to clearly distinguish "nature" and "convention".[57] The early Greek philosophers of the Milesian school, which was founded by Thales of Miletus and later continued by his successors Anaximander and Anaximenes, were the first to attempt to explain natural phenomena without relying on the supernatural.[58] The Pythagoreans developed a complex number philosophy[59]: 467–68  and contributed significantly to the development of mathematical science.[59]: 465  The theory of atoms was developed by the Greek philosopher Leucippus and his student Democritus.[60][61] Later, Epicurus would develop a full natural cosmology based on atomism, and would adopt a "canon" (ruler, standard) which established physical criteria or standards of scientific truth.[62] The Greek doctor Hippocrates established the tradition of systematic medical science[63][64] and is known as "The Father of Medicine".[65] A turning point in the history of early philosophical science was Socrates' example of applying philosophy to the study of human matters, including human nature, the nature of political communities, and human knowledge itself. The Socratic method as documented by Plato's dialogues is a dialectic method of hypothesis elimination: better hypotheses are found by steadily identifying and eliminating those that lead to contradictions. The Socratic method searches for general commonly-held truths that shape beliefs and scrutinizes them for consistency.[66] Socrates criticized the older type of study of physics as too purely speculative and lacking in self-criticism.[67] Aristotle in the 4th century BCE created a systematic program of teleological philosophy.[68] In the 3rd century BCE, Greek astronomer Aristarchus of Samos was the first to propose a heliocentric model of the universe, with the Sun at the center and all the planets orbiting it.[69] Aristarchus's model was widely rejected because it was believed to violate the laws of physics,[69] while Ptolemy's Almagest, which contains a geocentric description of the Solar System, was accepted through the early Renaissance instead.[70][71] The inventor and mathematician Archimedes of Syracuse made major contributions to the beginnings of calculus.[72] Pliny the Elder was a Roman writer and polymath, who wrote the seminal encyclopedia Natural History.[73][74][75] Positional notation for representing numbers likely emerged between the 3rd and 5th centuries CE along Indian trade routes. This numeral system made efficient arithmetic operations more accessible and would eventually become standard for mathematics worldwide.[76] Middle Ages Main article: History of science § Middle Ages Picture of a peacock on very old paper The first page of Vienna Dioscurides depicts a peacock, made in the 6th century Due to the collapse of the Western Roman Empire, the 5th century saw an intellectual decline and knowledge of Greek conceptions of the world deteriorated in Western Europe.[15]: 194  During the period, Latin encyclopedists such as Isidore of Seville preserved the majority of general ancient knowledge.[77] In contrast, because the Byzantine Empire resisted attacks from invaders, they were able to preserve and improve prior learning.[15]: 159  John Philoponus, a Byzantine scholar in the 500s, started to question Aristotle's teaching of physics, introducing the theory of impetus.[15]: 307, 311, 363, 402  His criticism served as an inspiration to medieval scholars and Galileo Galilei, who extensively cited his works ten centuries later.[15]: 307–308 [78] During late antiquity and the early Middle Ages, natural phenomena were mainly examined via the Aristotelian approach. The approach includes Aristotle's four causes: material, formal, moving, and final cause.[79] Many Greek classical texts were preserved by the Byzantine empire and Arabic translations were done by groups such as the Nestorians and the Monophysites. Under the Caliphate, these Arabic translations were later improved and developed by Arabic scientists.[80] By the 6th and 7th centuries, the neighboring Sassanid Empire established the medical Academy of Gondeshapur, which is considered by Greek, Syriac, and Persian physicians as the most important medical center of the ancient world.[81] The House of Wisdom was established in Abbasid-era Baghdad, Iraq,[82] where the Islamic study of Aristotelianism flourished[83] until the Mongol invasions in the 13th century. Ibn al-Haytham, better known as Alhazen, used controlled experiment in his optical study.[a][85][86] Avicenna's compilation of the Canon of Medicine, a medical encyclopedia, is considered to be one of the most important publications in medicine and was used until the 18th century.[87] By the eleventh century, most of Europe had become Christian,[15]: 204  and in 1088, the University of Bologna emerged as the first university in Europe.[88] As such, demand for Latin translation of ancient and scientific texts grew,[15]: 204  a major contributor to the Renaissance of the 12th century. Renaissance scholasticism in western Europe flourished, with experiments done by observing, describing, and classifying subjects in nature.[89] In the 13th century, medical teachers and students at Bologna began opening human bodies, leading to the first anatomy textbook based on human dissection by Mondino de Luzzi.[90] Renaissance Main articles: Scientific Revolution and Science in the Renaissance Drawing of planets' orbit around the Sun Drawing of the heliocentric model as proposed by the Copernicus's De revolutionibus orbium coelestium New developments in optics played a role in the inception of the Renaissance, both by challenging long-held metaphysical ideas on perception, as well as by contributing to the improvement and development of technology such as the camera obscura and the telescope. At the start of the Renaissance, Roger Bacon, Vitello, and John Peckham each built up a scholastic ontology upon a causal chain beginning with sensation, perception, and finally apperception of the individual and universal forms of Aristotle.[84]: Book I  A model of vision later known as perspectivism was exploited and studied by the artists of the Renaissance. This theory uses only three of Aristotle's four causes: formal, material, and final.[91] In the sixteenth century, Nicolaus Copernicus formulated a heliocentric model of the Solar System, stating that the planets revolve around the Sun, instead of the geocentric model where the planets and the Sun revolve around the Earth. This was based on a theorem that the orbital periods of the planets are longer as their orbs are farther from the center of motion, which he found not to agree with Ptolemy's model.[92] Johannes Kepler and others challenged the notion that the only function of the eye is perception, and shifted the main focus in optics from the eye to the propagation of light.[91][93] Kepler is best known, however, for improving Copernicus' heliocentric model through the discovery of Kepler's laws of planetary motion. Kepler did not reject Aristotelian metaphysics and described his work as a search for the Harmony of the Spheres.[94] Galileo had made significant contributions to astronomy, physics and engineering. However, he became persecuted after Pope Urban VIII sentenced him for writing about the heliocentric model.[95] The printing press was widely used to publish scholarly arguments, including some that disagreed widely with contemporary ideas of nature.[96] Francis Bacon and René Descartes published philosophical arguments in favor of a new type of non-Aristotelian science. Bacon emphasized the importance of experiment over contemplation, questioned the Aristotelian concepts of formal and final cause, promoted the idea that science should study the laws of nature and the improvement of all human life.[97] Descartes emphasized individual thought and argued that mathematics rather than geometry should be used to study nature.[98] Age of Enlightenment Main article: Science in the Age of Enlightenment see caption Title page of the 1687 first edition of Philosophiæ Naturalis Principia Mathematica by Isaac Newton At the start of the Age of Enlightenment, Isaac Newton formed the foundation of classical mechanics by his Philosophiæ Naturalis Principia Mathematica, greatly influencing future physicists.[99] Gottfried Wilhelm Leibniz incorporated terms from Aristotelian physics, now used in a new non-teleological way. This implied a shift in the view of objects: objects were now considered as having no innate goals. Leibniz assumed that different types of things all work according to the same general laws of nature, with no special formal or final causes.[100] During this time, the declared purpose and value of science became producing wealth and inventions that would improve human lives, in the materialistic sense of having more food, clothing, and other things. In Bacon's words, "the real and legitimate goal of sciences is the endowment of human life with new inventions and riches", and he discouraged scientists from pursuing intangible philosophical or spiritual ideas, which he believed contributed little to human happiness beyond "the fume of subtle, sublime or pleasing [speculation]".[101] Science during the Enlightenment was dominated by scientific societies and academies,[102] which had largely replaced universities as centers of scientific research and development. Societies and academies were the backbones of the maturation of the scientific profession. Another important development was the popularization of science among an increasingly literate population.[103] Enlightenment philosophers turned to a few of their scientific predecessors – Galileo, Kepler, Boyle, and Newton principally – as the guides to every physical and social field of the day.[104][105] The 18th century saw significant advancements in the practice of medicine[106] and physics;[107] the development of biological taxonomy by Carl Linnaeus;[108] a new understanding of magnetism and electricity;[109] and the maturation of chemistry as a discipline.[110] Ideas on human nature, society, and economics evolved during the Enlightenment. Hume and other Scottish Enlightenment thinkers developed A Treatise of Human Nature, which was expressed historically in works by authors including James Burnett, Adam Ferguson, John Millar and William Robertson, all of whom merged a scientific study of how humans behaved in ancient and primitive cultures with a strong awareness of the determining forces of modernity.[111] Modern sociology largely originated from this movement.[112] In 1776, Adam Smith published The Wealth of Nations, which is often considered the first work on modern economics.[113] 19th century Main article: 19th century in science Sketch of a map with captions The first diagram of an evolutionary tree made by Charles Darwin in 1837 During the nineteenth century, many distinguishing characteristics of contemporary modern science began to take shape. These included the transformation of the life and physical sciences, frequent use of precision instruments, emergence of terms such as "biologist", "physicist", "scientist", increased professionalization of those studying nature, scientists gained cultural authority over many dimensions of society, industrialization of numerous countries, thriving of popular science writings and emergence of science journals.[114] During the late 19th century, psychology emerged as a separate discipline from philosophy when Wilhelm Wundt founded the first laboratory for psychological research in 1879.[115] During the mid-19th century, Charles Darwin and Alfred Russel Wallace independently proposed the theory of evolution by natural selection in 1858, which explained how different plants and animals originated and evolved. Their theory was set out in detail in Darwin's book On the Origin of Species, published in 1859.[116] Separately, Gregor Mendel presented his paper, "Experiments on Plant Hybridization" in 1865,[117] which outlined the principles of biological inheritance, serving as the basis for modern genetics.[118] Early in the 19th century, John Dalton suggested the modern atomic theory, based on Democritus's original idea of indivisible particles called atoms.[119] The laws of conservation of energy, conservation of momentum and conservation of mass suggested a highly stable universe where there could be little loss of resources. However, with the advent of the steam engine and the industrial revolution there was an increased understanding that not all forms of energy have the same energy qualities, the ease of conversion to useful work or to another form of energy.[120] This realization led to the development of the laws of thermodynamics, in which the free energy of the universe is seen as constantly declining: the entropy of a closed universe increases over time.[b] The electromagnetic theory was established in the 19th century by the works of Hans Christian Ørsted, André-Marie Ampère, Michael Faraday, James Clerk Maxwell, Oliver Heaviside, and Heinrich Hertz. The new theory raised questions that could not easily be answered using Newton's framework. The discovery of X-rays inspired the discovery of radioactivity by Henri Becquerel and Marie Curie in 1896,[123] Marie Curie then became the first person to win two Nobel prizes.[124] In the next year came the discovery of the first subatomic particle, the electron.[125] 20th century Main article: 20th century in science Graph showing lower ozone concentration at the South Pole A computer graph of the ozone hole made in 1987 using data from a space telescope In the first half of the century, the development of antibiotics and artificial fertilizers improved human living standards globally.[126][127] Harmful environmental issues such as ozone depletion, ocean acidification, eutrophication and climate change came to the public's attention and caused the onset of environmental studies.[128] During this period, scientific experimentation became increasingly larger in scale and funding.[129] The extensive technological innovation stimulated by World War I, World War II, and the Cold War led to competitions between global powers, such as the Space Race and nuclear arms race.[130][131] Substantial international collaborations were also made, despite armed conflicts.[132] In the late 20th century, active recruitment of women and elimination of sex discrimination greatly increased the number of women scientists, but large gender disparities remained in some fields.[133] The discovery of the cosmic microwave background in 1964[134] led to a rejection of the steady-state model of the universe in favor of the Big Bang theory of Georges Lemaître.[135] The century saw fundamental changes within science disciplines. Evolution became a unified theory in the early 20th-century when the modern synthesis reconciled Darwinian evolution with classical genetics.[136] Albert Einstein's theory of relativity and the development of quantum mechanics complement classical mechanics to describe physics in extreme length, time and gravity.[137][138] Widespread use of integrated circuits in the last quarter of the 20th century combined with communications satellites led to a revolution in information technology and the rise of the global internet and mobile computing, including smartphones. The need for mass systematization of long, intertwined causal chains and large amounts of data led to the rise of the fields of systems theory and computer-assisted scientific modeling.[139] 21st century Main article: 21st century § Science and technology Four predicted image of M87* black hole made by separate teams in the Event Horizon Telescope collaboration. The Human Genome Project was completed in 2003 by identifying and mapping all of the genes of the human genome.[140] The first induced pluripotent human stem cells were made in 2006, allowing adult cells to be transformed into stem cells and turn to any cell type found in the body.[141] With the affirmation of the Higgs boson discovery in 2013, the last particle predicted by the Standard Model of particle physics was found.[142] In 2015, gravitational waves, predicted by general relativity a century before, were first observed.[143][144] In 2019, the international collaboration Event Horizon Telescope presented the first direct image of a black hole's accretion disk.[145] Branches Main article: Branches of science Modern science is commonly divided into three major branches: natural science, social science, and formal science.[3] Each of these branches comprises various specialized yet overlapping scientific disciplines that often possess their own nomenclature and expertise.[146] Both natural and social sciences are empirical sciences,[147] as their knowledge is based on empirical observations and is capable of being tested for its validity by other researchers working under the same conditions.[148] Natural science Natural science is the study of the physical world. It can be divided into two main branches: life science and physical science. These two branches may be further divided into more specialized disciplines. For example, physical science can be subdivided into physics, chemistry, astronomy, and earth science. Modern natural science is the successor to the natural philosophy that began in Ancient Greece. Galileo, Descartes, Bacon, and Newton debated the benefits of using approaches which were more mathematical and more experimental in a methodical way. Still, philosophical perspectives, conjectures, and presuppositions, often overlooked, remain necessary in natural science.[149] Systematic data collection, including discovery science, succeeded natural history, which emerged in the 16th century by describing and classifying plants, animals, minerals, and so on.[150] Today, "natural history" suggests observational descriptions aimed at popular audiences.[151] Social science Two curve crossing over at a point, forming a X shape Supply and demand curve in economics, crossing over at the optimal equilibrium Social science is the study of human behavior and functioning of societies.[4][5] It has many disciplines that include, but are not limited to anthropology, economics, history, human geography, political science, psychology, and sociology.[4] In the social sciences, there are many competing theoretical perspectives, many of which are extended through competing research programs such as the functionalists, conflict theorists, and interactionists in sociology.[4] Due to the limitations of conducting controlled experiments involving large groups of individuals or complex situations, social scientists may adopt other research methods such as the historical method, case studies, and cross-cultural studies. Moreover, if quantitative information is available, social scientists may rely on statistical approaches to better understand social relationships and processes.[4] Formal science Formal science is an area of study that generates knowledge using formal systems.[152][6][7] A formal system is an abstract structure used for inferring theorems from axioms according to a set of rules.[153] It includes mathematics,[154][155] systems theory, and theoretical computer science. The formal sciences share similarities with the other two branches by relying on objective, careful, and systematic study of an area of knowledge. They are, however, different from the empirical sciences as they rely exclusively on deductive reasoning, without the need for empirical evidence, to verify their abstract concepts.[11][156][148] The formal sciences are therefore a priori disciplines and because of this, there is disagreement on whether they constitute a science.[8][157] Nevertheless, the formal sciences play an important role in the empirical sciences. Calculus, for example, was initially invented to understand motion in physics.[158] Natural and social sciences that rely heavily on mathematical applications include mathematical physics,[159] chemistry,[160] biology,[161] finance,[162] and economics.[163] Applied science Applied science is the use of the scientific method and knowledge to attain practical goals and includes a broad range of disciplines such as engineering and medicine.[164][14] Engineering is the use of scientific principles to invent, design and build machines, structures and technologies.[165] Science may contribute to the development of new technologies.[166] Medicine is the practice of caring for patients by maintaining and restoring health through the prevention, diagnosis, and treatment of injury or disease.[167][168] The applied sciences are often contrasted with the basic sciences, which are focused on advancing scientific theories and laws that explain and predict events in the natural world.[169][170] Computational science applies computing power to simulate real-world situations, enabling a better understanding of scientific problems than formal mathematics alone can achieve. The use of machine learning and artificial intelligence is becoming a central feature of computational contributions to science for example in agent-based computational economics, random forests, topic modeling and various forms of prediction. However, machines alone rarely advance knowledge as they require human guidance and capacity to reason; and they can introduce bias against certain social groups or sometimes underperform against humans.[171][172] Interdisciplinary science Interdisciplinary science involves the combination of two or more disciplines into one,[173] such as bioinformatics, a combination of biology and computer science[174] or cognitive sciences. The concept has existed since the ancient Greek and it became popular again in the 20th century.[175] Scientific research Scientific research can be labeled as either basic or applied research. Basic research is the search for knowledge and applied research is the search for solutions to practical problems using this knowledge. Most understanding comes from basic research, though sometimes applied research targets specific practical problems. This leads to technological advances that were not previously imaginable.[176] Scientific method 6 steps of the scientific method in a loop A diagram variant of scientific method represented as an ongoing process[broken anchor] Scientific research involves using the scientific method, which seeks to objectively explain the events of nature in a reproducible way.[177] Scientists usually take for granted a set of basic assumptions that are needed to justify the scientific method: there is an objective reality shared by all rational observers; this objective reality is governed by natural laws; these laws were discovered by means of systematic observation and experimentation.[2] Mathematics is essential in the formation of hypotheses, theories, and laws, because it is used extensively in quantitative modeling, observing, and collecting measurements.[178] Statistics is used to summarize and analyze data, which allows scientists to assess the reliability of experimental results.[179] In the scientific method, an explanatory thought experiment or hypothesis is put forward as an explanation using parsimony principles and is expected to seek consilience – fitting with other accepted facts related to an observation or scientific question.[180] This tentative explanation is used to make falsifiable predictions, which are typically posted before being tested by experimentation. Disproof of a prediction is evidence of progress.[177]: 4–5 [181] Experimentation is especially important in science to help establish causal relationships to avoid the correlation fallacy, though in some sciences such as astronomy or geology, a predicted observation might be more appropriate.[182] When a hypothesis proves unsatisfactory, it is modified or discarded.[183] If the hypothesis survived testing, it may become adopted into the framework of a scientific theory, a validly reasoned, self-consistent model or framework for describing the behavior of certain natural events. A theory typically describes the behavior of much broader sets of observations than a hypothesis; commonly, a large number of hypotheses can be logically bound together by a single theory. Thus a theory is a hypothesis explaining various other hypotheses. In that vein, theories are formulated according to most of the same scientific principles as hypotheses. Scientists may generate a model, an attempt to describe or depict an observation in terms of a logical, physical or mathematical representation and to generate new hypotheses that can be tested by experimentation.[184] While performing experiments to test hypotheses, scientists may have a preference for one outcome over another.[185][186] Eliminating the bias can be achieved by transparency, careful experimental design, and a thorough peer review process of the experimental results and conclusions.[187][188] After the results of an experiment are announced or published, it is normal practice for independent researchers to double-check how the research was performed, and to follow up by performing similar experiments to determine how dependable the results might be.[189] Taken in its entirety, the scientific method allows for highly creative problem solving while minimizing the effects of subjective and confirmation bias.[190] Intersubjective verifiability, the ability to reach a consensus and reproduce results, is fundamental to the creation of all scientific knowledge.[191] Scientific literature Main articles: Scientific literature and Lists of important publications in science Decorated "NATURE" as title, with scientific text below Cover of the first issue of Nature, November 4, 1869 Scientific research is published in a range of literature.[192] Scientific journals communicate and document the results of research carried out in universities and various other research institutions, serving as an archival record of science. The first scientific journals, Journal des sçavans followed by Philosophical Transactions, began publication in 1665. Since that time the total number of active periodicals has steadily increased. In 1981, one estimate for the number of scientific and technical journals in publication was 11,500.[193] Most scientific journals cover a single scientific field and publish the research within that field; the research is normally expressed in the form of a scientific paper. Science has become so pervasive in modern societies that it is considered necessary to communicate the achievements, news, and ambitions of scientists to a wider population.[194] Challenges The replication crisis is an ongoing methodological crisis that affects parts of the social and life sciences. In subsequent investigations, the results of many scientific studies are proven to be unrepeatable.[195] The crisis has long-standing roots; the phrase was coined in the early 2010s[196] as part of a growing awareness of the problem. The replication crisis represents an important body of research in metascience, which aims to improve the quality of all scientific research while reducing waste.[197] An area of study or speculation that masquerades as science in an attempt to claim a legitimacy that it would not otherwise be able to achieve is sometimes referred to as pseudoscience, fringe science, or junk science.[198][199] Physicist Richard Feynman coined the term "cargo cult science" for cases in which researchers believe and at a glance looks like they are doing science, but lack the honesty allowing their results to be rigorously evaluated.[200] Various types of commercial advertising, ranging from hype to fraud, may fall into these categories. Science has been described as "the most important tool" for separating valid claims from invalid ones.[201] There can also be an element of political or ideological bias on all sides of scientific debates. Sometimes, research may be characterized as "bad science," research that may be well-intended but is incorrect, obsolete, incomplete, or over-simplified expositions of scientific ideas. The term "scientific misconduct" refers to situations such as where researchers have intentionally misrepresented their published data or have purposely given credit for a discovery to the wrong person.[202] Philosophy of science Depiction of epicycles, where a planet orbit is going around in a bigger orbit For Kuhn, the addition of epicycles in Ptolemaic astronomy was "normal science" within a paradigm, whereas the Copernican Revolution was a paradigm shift There are different schools of thought in the philosophy of science. The most popular position is empiricism, which holds that knowledge is created by a process involving observation; scientific theories generalize observations.[203] Empiricism generally encompasses inductivism, a position that explains how general theories can be made from the finite amount of empirical evidence available. Many versions of empiricism exist, with the predominant ones being Bayesianism and the hypothetico-deductive method.[204][203] Empiricism has stood in contrast to rationalism, the position originally associated with Descartes, which holds that knowledge is created by the human intellect, not by observation.[205] Critical rationalism is a contrasting 20th-century approach to science, first defined by Austrian-British philosopher Karl Popper. Popper rejected the way that empiricism describes the connection between theory and observation. He claimed that theories are not generated by observation, but that observation is made in the light of theories: that the only way theory A can be affected by observation is after theory A were to conflict with observation, but theory B were to survive the observation.[206] Popper proposed replacing verifiability with falsifiability as the landmark of scientific theories, replacing induction with falsification as the empirical method.[206] Popper further claimed that there is actually only one universal method, not specific to science: the negative method of criticism, trial and error,[207] covering all products of the human mind, including science, mathematics, philosophy, and art.[208] Another approach, instrumentalism, emphasizes the utility of theories as instruments for explaining and predicting phenomena. It views scientific theories as black boxes with only their input (initial conditions) and output (predictions) being relevant. Consequences, theoretical entities, and logical structure are claimed to be something that should be ignored.[209] Close to instrumentalism is constructive empiricism, according to which the main criterion for the success of a scientific theory is whether what it says about observable entities is true.[210] Thomas Kuhn argued that the process of observation and evaluation takes place within a paradigm, a logically consistent "portrait" of the world that is consistent with observations made from its framing. He characterized normal science as the process of observation and "puzzle solving" which takes place within a paradigm, whereas revolutionary science occurs when one paradigm overtakes another in a paradigm shift.[211] Each paradigm has its own distinct questions, aims, and interpretations. The choice between paradigms involves setting two or more "portraits" against the world and deciding which likeness is most promising. A paradigm shift occurs when a significant number of observational anomalies arise in the old paradigm and a new paradigm makes sense of them. That is, the choice of a new paradigm is based on observations, even though those observations are made against the background of the old paradigm. For Kuhn, acceptance or rejection of a paradigm is a social process as much as a logical process. Kuhn's position, however, is not one of relativism.[212] Finally, another approach often cited in debates of scientific skepticism against controversial movements like "creation science" is methodological naturalism. Naturalists maintain that a difference should be made between natural and supernatural, and science should be restricted to natural explanations.[213] Methodological naturalism maintains that science requires strict adherence to empirical study and independent verification.[214] Scientific community The scientific community is a network of interacting scientists who conducts scientific research. The community consists of smaller groups working in scientific fields. By having peer review, through discussion and debate within journals and conferences, scientists maintain the quality of research methodology and objectivity when interpreting results.[215] Scientists Portrait of a middle-aged woman Marie Curie was the first person to be awarded two Nobel Prizes: Physics in 1903 and Chemistry in 1911[124] Scientists are individuals who conduct scientific research to advance knowledge in an area of interest.[216][217] In modern times, many professional scientists are trained in an academic setting and upon completion, attain an academic degree, with the highest degree being a doctorate such as a Doctor of Philosophy or PhD.[218] Many scientists pursue careers in various sectors of the economy such as academia, industry, government, and nonprofit organizations.[219][220][221] Scientists exhibit a strong curiosity about reality and a desire to apply scientific knowledge for the benefit of health, nations, the environment, or industries. Other motivations include recognition by their peers and prestige. In modern times, many scientists have advanced degrees in an area of science and pursue careers in various sectors of the economy such as academia, industry, government, and nonprofit environments.[222] [223][224] Science has historically been a male-dominated field, with notable exceptions. Women in science faced considerable discrimination in science, much as they did in other areas of male-dominated societies. For example, women were frequently being passed over for job opportunities and denied credit for their work.[225] The achievements of women in science have been attributed to the defiance of their traditional role as laborers within the domestic sphere.[226] Learned societies Picture of scientists in 200th anniversary of the Prussian Academy of Sciences, 1900 Learned societies for the communication and promotion of scientific thought and experimentation have existed since the Renaissance.[227] Many scientists belong to a learned society that promotes their respective scientific discipline, profession, or group of related disciplines.[228] Membership may either be open to all, require possession of scientific credentials, or conferred by election.[229] Most scientific societies are non-profit organizations,[230] and many are professional associations. Their activities typically include holding regular conferences for the presentation and discussion of new research results and publishing or sponsoring academic journals in their discipline. Some societies act as professional bodies, regulating the activities of their members in the public interest, or the collective interest of the membership. The professionalization of science, begun in the 19th century, was partly enabled by the creation of national distinguished academies of sciences such as the Italian Accademia dei Lincei in 1603,[231] the British Royal Society in 1660,[232] the French Academy of Sciences in 1666,[233] the American National Academy of Sciences in 1863,[234] the German Kaiser Wilhelm Society in 1911,[235] and the Chinese Academy of Sciences in 1949.[236] International scientific organizations, such as the International Science Council, are devoted to international cooperation for science advancement.[237] Awards Science awards are usually given to individuals or organizations that have made significant contributions to a discipline. They are often given by prestigious institutions, thus it is considered a great honor for a scientist receiving them. Since the early Renaissance, scientists are often awarded medals, money, and titles. The Nobel Prize, a widely regarded prestigious award, is awarded annually to those who have achieved scientific advances in the fields of medicine, physics, and chemistry.[238] Society "Science and society" redirects here. Not to be confused with Science & Society or Sociology of scientific knowledge. Funding and policies see caption Budget of NASA as percentage of United States federal budget, peaking at 4.4% in 1966 and slowly declining since Scientific research is often funded through a competitive process in which potential research projects are evaluated and only the most promising receive funding. Such processes, which are run by government, corporations, or foundations, allocate scarce funds. Total research funding in most developed countries is between 1.5% and 3% of GDP.[239] In the OECD, around two-thirds of research and development in scientific and technical fields is carried out by industry, and 20% and 10% respectively by universities and government. The government funding proportion in certain fields is higher, and it dominates research in social science and humanities. In the lesser-developed nations, government provides the bulk of the funds for their basic scientific research.[240] Many governments have dedicated agencies to support scientific research, such as the National Science Foundation in the United States,[241] the National Scientific and Technical Research Council in Argentina,[242] Commonwealth Scientific and Industrial Research Organization in Australia,[243] National Centre for Scientific Research in France,[244] the Max Planck Society in Germany,[245] and National Research Council in Spain.[246] In commercial research and development, all but the most research-oriented corporations focus more heavily on near-term commercialization possibilities rather than research driven by curiosity.[247] Science policy is concerned with policies that affect the conduct of the scientific enterprise, including research funding, often in pursuance of other national policy goals such as technological innovation to promote commercial product development, weapons development, health care, and environmental monitoring. Science policy sometimes refers to the act of applying scientific knowledge and consensus to the development of public policies. In accordance with public policy being concerned about the well-being of its citizens, science policy's goal is to consider how science and technology can best serve the public.[248] Public policy can directly affect the funding of capital equipment and intellectual infrastructure for industrial research by providing tax incentives to those organizations that fund research.[194] Education and awareness Main articles: Public awareness of science and Science journalism Dinosaur exhibit in the Houston Museum of Natural Science Science education for the general public is embedded in the school curriculum, and is supplemented by online pedagogical content (for example, YouTube and Khan Academy), museums, and science magazines and blogs. Scientific literacy is chiefly concerned with an understanding of the scientific method, units and methods of measurement, empiricism, a basic understanding of statistics (correlations, qualitative versus quantitative observations, aggregate statistics), as well as a basic understanding of core scientific fields, such as physics, chemistry, biology, ecology, geology and computation. As a student advances into higher stages of formal education, the curriculum becomes more in depth. Traditional subjects usually included in the curriculum are natural and formal sciences, although recent movements include social and applied science as well.[249] The mass media face pressures that can prevent them from accurately depicting competing scientific claims in terms of their credibility within the scientific community as a whole. Determining how much weight to give different sides in a scientific debate may require considerable expertise regarding the matter.[250] Few journalists have real scientific knowledge, and even beat reporters who are knowledgeable about certain scientific issues may be ignorant about other scientific issues that they are suddenly asked to cover.[251][252] Science magazines such as New Scientist, Science & Vie, and Scientific American cater to the needs of a much wider readership and provide a non-technical summary of popular areas of research, including notable discoveries and advances in certain fields of research.[253] Science fiction genre, primarily speculative fiction, can transmit the ideas and methods of science to the general public.[254] Recent efforts to intensify or develop links between science and non-scientific disciplines, such as literature or poetry, include the Creative Writing Science resource developed through the Royal Literary Fund.[255] Anti-science attitudes Main article: Antiscience While the scientific method is broadly accepted in the scientific community, some fractions of society reject certain scientific positions or are skeptical about science. Examples are the common notion that COVID-19 is not a major health threat to the US (held by 39% of Americans in August 2021)[256] or the belief that climate change is not a major threat to the US (also held by 40% of Americans, in late 2019 and early 2020).[257] Psychologists have pointed to four factors driving rejection of scientific results:[258] Scientific authorities are sometimes seen as inexpert, untrustworthy, or biased. Some marginalized social groups hold anti-science attitudes, in part because these groups have often been exploited in unethical experiments.[259] Messages from scientists may contradict deeply-held existing beliefs or morals. The delivery of a scientific message may not be appropriately targeted to a recipient's learning style. Anti-science attitudes seem to be often caused by fear of rejection in social groups. For instance, climate change is perceived as a threat by only 22% of Americans on the right side of the political spectrum, but by 85% on the left.[260] That is, if someone on the left would not consider climate change as a threat, this person may face contempt and be rejected in that social group. In fact, people may rather deny a scientifically accepted fact than lose or jeopardize their social status.[261] Politics Result in bar graph of two questions ("Is global warming occurring?" and "Are oil/gas companies responsible?"), showing large discrepancies between American Democrats and Republicans Public opinion on global warming in the United States by political party[262] Attitudes towards science are often determined by political opinions and goals. Government, business and advocacy groups have been known to use legal and economic pressure to influence scientific researchers. Many factors can act as facets of the politicization of science such as anti-intellectualism, perceived threats to religious beliefs, and fear for business interests.[263] Politicization of science is usually accomplished when scientific information is presented in a way that emphasizes the uncertainty associated with the scientific evidence.[264] Tactics such as shifting conversation, failing to acknowledge facts, and capitalizing on doubt of scientific consensus have been used to gain more attention for views that have been undermined by scientific evidence.[265] Examples of issues that have involved the politicization of science include the global warming controversy, health effects of pesticides, and health effects of tobacco.[265][266] See also Criticism of science List of scientific occupations List of years in science Notes Ibn al-Haytham's Book of Optics Book I, [6.54]. pages 372 and 408 disputed Claudius Ptolemy's extramission theory of vision; "Hence, the extramission of [visual] rays is superfluous and useless". —A.Mark Smith's translation of the Latin version of Ibn al-Haytham.[84]: Book I, [6.54]. pp. 372, 408  Whether the universe is closed or open, or the shape of the universe, is an open question. The 2nd law of thermodynamics,[120]: 9 [121] and the 3rd law of thermodynamics[122] imply the heat death of the universe if the universe is a closed system, but not necessarily for an expanding universe. References Wilson, E.O. (1999). "The natural sciences". Consilience: The Unity of Knowledge (Reprint ed.). New York: Vintage. pp. 49–71. ISBN 978-0-679-76867-8. Heilbron, J.L.; et al. (2003). "Preface". The Oxford Companion to the History of Modern Science. New York: Oxford University Press. pp. vii–x. ISBN 978-0-19-511229-0. ...modern science is a discovery as well as an invention. It was a discovery that nature generally acts regularly enough to be described by laws and even by mathematics; and required invention to devise the techniques, abstractions, apparatus, and organization for exhibiting the regularities and securing their law-like descriptions. Cohen, Eliel (2021). "The boundary lens: theorising academic actitity". The University and its Boundaries: Thriving or Surviving in the 21st Century. New York: Routledge. pp. 14–41. ISBN 978-0-367-56298-4. Archived from the original on May 5, 2021. Retrieved May 4, 2021. Colander, David C.; Hunt, Elgin F. (2019). "Social science and its methods". Social Science: An Introduction to the Study of Society (17th ed.). New York, NY: Routledge. pp. 1–22. Nisbet, Robert A.; Greenfeld, Liah (October 16, 2020). "Social Science". Encyclopedia Britannica. Encyclopædia Britannica, Inc. Archived from the original on February 2, 2022. Retrieved May 9, 2021. Löwe, Benedikt (2002). "The formal sciences: their scope, their foundations, and their unity". Synthese. 133 (1/2): 5–11. doi:10.1023/A:1020887832028. S2CID 9272212. Rucker, Rudy (2019). "Robots and souls". Infinity and the Mind: The Science and Philosophy of the Infinite (Reprint ed.). Princeton, New Jersey: Princeton University Press. pp. 157–188. ISBN 978-0-691-19138-6. Archived from the original on February 26, 2021. Retrieved May 11, 2021. Bishop, Alan (1991). "Environmental activities and mathematical culture". Mathematical Enculturation: A Cultural Perspective on Mathematics Education. Norwell, Massachusetts: Kluwer Academic Publishers. pp. 20–59. ISBN 978-0-7923-1270-3. Archived from the original on December 25, 2020. Retrieved March 24, 2018. Nickles, Thomas (2013). "The Problem of Demarcation". Philosophy of Pseudoscience: Reconsidering the Demarcation Problem. Chicago: The University of Chicago Press. p. 104. Bunge, Mario (1998). "The Scientific Approach". Philosophy of Science. Vol. 1, From Problem to Theory (revised ed.). New York: Routledge. pp. 3–50. ISBN 978-0-7658-0413-6. Fetzer, James H. (2013). "Computer reliability and public policy: Limits of knowledge of computer-based systems". Computers and Cognition: Why Minds are not Machines. Newcastle, United Kingdom: Kluwer Academic Publishers. pp. 271–308. ISBN 978-1-4438-1946-6. Fischer, M.R.; Fabry, G (2014). "Thinking and acting scientifically: Indispensable basis of medical education". GMS Zeitschrift für Medizinische Ausbildung. 31 (2): Doc24. doi:10.3205/zma000916. PMC 4027809. PMID 24872859. Sinclair, Marius (1993). "On the Differences between the Engineering and Scientific Methods". The International Journal of Engineering Education. Archived from the original on November 15, 2017. Retrieved September 7, 2018. Bunge, M (1966). "Technology as Applied Science". In Rapp, F. (ed.). Contributions to a Philosophy of Technology. Dordrecht, Netherlands: Springer. pp. 19–39. doi:10.1007/978-94-010-2182-1_2. ISBN 978-94-010-2184-5. S2CID 110332727. Lindberg, David C. (2007). The beginnings of Western science: the European Scientific tradition in philosophical, religious, and institutional context (2nd ed.). University of Chicago Press. ISBN 978-0226482057. Grant, Edward (2007). "Ancient Egypt to Plato". A History of Natural Philosophy: From the Ancient World to the Nineteenth Century. New York: Cambridge University Press. pp. 1–26. ISBN 978-0-521-68957-1. Building Bridges Among the BRICs, p. 125, Robert Crane, Springer, 2014 Keay, John (2000). India: A history. Atlantic Monthly Press. p. 132. ISBN 978-0-87113-800-2. The great era of all that is deemed classical in Indian literature, art and science was now dawning. It was this crescendo of creativity and scholarship, as much as ... political achievements of the Guptas, which would make their age so golden. Lindberg, David C. (2007). "Islamic science". The beginnings of Western science: the European Scientific tradition in philosophical, religious, and institutional context (2nd ed.). Chicago: University of Chicago Press. pp. 163–92. ISBN 978-0-226-48205-7. Lindberg, David C. (2007). "The revival of learning in the West". The beginnings of Western science: the European Scientific tradition in philosophical, religious, and institutional context (2nd ed.). Chicago: University of Chicago Press. pp. 193–224. ISBN 978-0-226-48205-7. Lindberg, David C. (2007). "The recovery and assimilation of Greek and Islamic science". The beginnings of Western science: the European Scientific tradition in philosophical, religious, and institutional context (2nd ed.). Chicago: University of Chicago Press. pp. 225–53. ISBN 978-0-226-48205-7. Sease, Virginia; Schmidt-Brabant, Manfrid. Thinkers, Saints, Heretics: Spiritual Paths of the Middle Ages. 2007. Pages 80-81. Retrieved 6 Oct. 2023 Principe, Lawrence M. (2011). "Introduction". Scientific Revolution: A Very Short Introduction. New York: Oxford University Press. pp. 1–3. ISBN 978-0-19-956741-6. Lindberg, David C. (2007). "The legacy of ancient and medieval science". The beginnings of Western science: the European Scientific tradition in philosophical, religious, and institutional context (2nd ed.). Chicago: University of Chicago Press. pp. 357–368. ISBN 978-0-226-48205-7. Grant, Edward (2007). "Transformation of medieval natural philosophy from the early period modern period to the end of the nineteenth century". A History of Natural Philosophy: From the Ancient World to the Nineteenth Century. New York: Cambridge University Press. pp. 274–322. ISBN 978-0-521-68957-1. Cahan, David, ed. (2003). From Natural Philosophy to the Sciences: Writing the History of Nineteenth-Century Science. Chicago: University of Chicago Press. ISBN 978-0-226-08928-7. Lightman, Bernard (2011). "13. Science and the Public". In Shank, Michael; Numbers, Ronald; Harrison, Peter (eds.). Wrestling with Nature: From Omens to Science. Chicago: University of Chicago Press. p. 367. ISBN 978-0-226-31783-0. Harrison, Peter (2015). The Territories of Science and Religion. Chicago: University of Chicago Press. pp. 164–165. ISBN 978-0-226-18451-7. The changing character of those engaged in scientific endeavors was matched by a new nomenclature for their endeavors. The most conspicuous marker of this change was the replacement of "natural philosophy" by "natural science". In 1800 few had spoken of the "natural sciences" but by 1880, this expression had overtaken the traditional label "natural philosophy". The persistence of "natural philosophy" in the twentieth century is owing largely to historical references to a past practice (see figure 11). As should now be apparent, this was not simply the substitution of one term by another, but involved the jettisoning of a range of personal qualities relating to the conduct of philosophy and the living of the philosophical life. MacRitchie, Finlay (2011). "Introduction". Scientific Research as a Career. New York: Routledge. pp. 1–6. ISBN 978-1-4398-6965-9. Archived from the original on May 5, 2021. Retrieved May 5, 2021. Marder, Michael P. (2011). "Curiosity and research". Research Methods for Science. New York: Cambridge University Press. pp. 1–17. ISBN 978-0-521-14584-8. Archived from the original on May 5, 2021. Retrieved May 5, 2021. de Ridder, Jeroen (2020). "How many scientists does it take to have knowledge?". In McCain, Kevin; Kampourakis, Kostas (eds.). What is Scientific Knowledge? An Introduction to Contemporary Epistemology of Science. New York: Routledge. pp. 3–17. ISBN 978-1-138-57016-0. Archived from the original on May 5, 2021. Retrieved May 5, 2021. Lindberg, David C. (2007). "Islamic science". The beginnings of Western science: the European Scientific tradition in philosophical, religious, and institutional context (2nd ed.). Chicago: University of Chicago Press. pp. 163–192. ISBN 978-0-226-48205-7. Szycher, Michael (2016). "Establishing your dream team". Commercialization Secrets for Scientists and Engineers. New York: Routledge. pp. 159–176. ISBN 978-1-138-40741-1. Archived from the original on August 18, 2021. Retrieved May 5, 2021. "science". Merriam-Webster Online Dictionary. Merriam-Webster, Inc. Archived from the original on September 1, 2019. Retrieved October 16, 2011. Vaan, Michiel de (2008). "sciō". Etymological Dictionary of Latin and the other Italic Languages. Indo-European Etymological Dictionary. p. 545. ISBN 978-90-04-16797-1. Cahan, David (2003). From natural philosophy to the sciences : writing the history of nineteenth-century science. Chicago: University of Chicago Press. pp. 3–15. ISBN 0-226-08927-4. OCLC 51330464. Archived from the original on May 31, 2022. Retrieved May 31, 2022. Ross, Sydney (1962). "Scientist: The story of a word". Annals of Science. 18 (2): 65–85. doi:10.1080/00033796200202722. "scientist". Oxford English Dictionary (Online ed.). Oxford University Press. (Subscription or participating institution membership required.) Carruthers, Peter (May 2, 2002). Carruthers, Peter; Stich, Stephen; Siegal, Michael (eds.). "The roots of scientific reasoning: infancy, modularity and the art of tracking". The Cognitive Basis of Science. Cambridge University Press. pp. 73–96. doi:10.1017/cbo9780511613517.005. ISBN 978-0-521-81229-0. Lombard, Marlize; Gärdenfors, Peter (2017). "Tracking the Evolution of Causal Cognition in Humans". Journal of Anthropological Sciences. 95 (95): 219–234. doi:10.4436/JASS.95006. ISSN 1827-4765. PMID 28489015. Graeber, David; Wengrow, David (2021). The Dawn of Everything. p. 248. Budd, Paul; Taylor, Timothy (1995). "The Faerie Smith Meets the Bronze Industry: Magic Versus Science in the Interpretation of Prehistoric Metal-Making". World Archaeology. 27 (1): 133–143. doi:10.1080/00438243.1995.9980297. JSTOR 124782. Tuomela, Raimo (1987). "Science, Protoscience, and Pseudoscience". In Pitt, J.C.; Pera, M. (eds.). Rational Changes in Science. Boston Studies in the Philosophy of Science. Vol. 98. Dordrecht: Springer. pp. 83–101. doi:10.1007/978-94-009-3779-6_4. ISBN 978-94-010-8181-8. Smith, Pamela H. (2009). "Science on the Move: Recent Trends in the History of Early Modern Science". Renaissance Quarterly. 62 (2): 345–375. doi:10.1086/599864. PMID 19750597. S2CID 43643053. Fleck, Robert (March 2021). "Fundamental Themes in Physics from the History of Art". Physics in Perspective. 23 (1): 25–48. Bibcode:2021PhP....23...25F. doi:10.1007/s00016-020-00269-7. ISSN 1422-6944. S2CID 253597172. Scott, Colin (2011). "Science for the West, Myth for the Rest?". In Harding, Sandra (ed.). The Postcolonial Science and Technology Studies Reader. Durham: Duke University Press. p. 175. doi:10.2307/j.ctv11g96cc.16. ISBN 978-0-8223-4936-5. OCLC 700406626. Dear, Peter (2012). "Historiography of Not-So-Recent Science". History of Science. 50 (2): 197–211. doi:10.1177/007327531205000203. S2CID 141599452. Rochberg, Francesca (2011). "Ch.1 Natural Knowledge in Ancient Mesopotamia". In Shank, Michael; Numbers, Ronald; Harrison, Peter (eds.). Wrestling with Nature: From Omens to Science. Chicago: University of Chicago Press. p. 9. ISBN 978-0-226-31783-0. Krebs, Robert E. (2004). Groundbreaking Scientific Experiments, Inventions, and Discoveries of the Middle Ages and the Renaissance. Greenwood Publishing Group. p. 127. ISBN 978-0313324338. Erlich, Ḥaggai; Gershoni, Israel (2000). The Nile: Histories, Cultures, Myths. Lynne Rienner Publishers. pp. 80–81. ISBN 978-1-55587-672-2. Archived from the original on May 31, 2022. Retrieved January 9, 2020. The Nile occupied an important position in Egyptian culture; it influenced the development of mathematics, geography, and the calendar; Egyptian geometry advanced due to the practice of land measurement "because the overflow of the Nile caused the boundary of each person's land to disappear." "Telling Time in Ancient Egypt". The Met's Heilbrunn Timeline of Art History. Archived from the original on March 3, 2022. Retrieved May 27, 2022. McIntosh, Jane R. (2005). Ancient Mesopotamia: New Perspectives. Santa Barbara, California, Denver, Colorado, and Oxford, England: ABC-CLIO. pp. 273–76. ISBN 978-1-57607-966-9. Archived from the original on February 5, 2021. Retrieved October 20, 2020. Aaboe, Asger (May 2, 1974). "Scientific Astronomy in Antiquity". Philosophical Transactions of the Royal Society. 276 (1257): 21–42. Bibcode:1974RSPTA.276...21A. doi:10.1098/rsta.1974.0007. JSTOR 74272. S2CID 122508567. Biggs, R D. (2005). "Medicine, Surgery, and Public Health in Ancient Mesopotamia". Journal of Assyrian Academic Studies. 19 (1): 7–18. Lehoux, Daryn (2011). "2. Natural Knowledge in the Classical World". In Shank, Michael; Numbers, Ronald; Harrison, Peter (eds.). Wrestling with Nature: From Omens to Science. Chicago: University of Chicago Press. p. 39. ISBN 978-0-226-31783-0. An account of the pre-Socratic use of the concept of φύσις may be found in Naddaf, Gerard (2006). The Greek Concept of Nature. SUNY Press, and in Ducarme, Frédéric; Couvet, Denis (2020). "What does 'nature' mean?" (PDF). Palgrave Communications. 6 (14). Springer Nature. doi:10.1057/s41599-020-0390-y. The word φύσις, while first used in connection with a plant in Homer, occurs early in Greek philosophy, and in several senses. Generally, these senses match rather well the current senses in which the English word nature is used, as confirmed by Guthrie, W.K.C. Presocratic Tradition from Parmenides to Democritus (volume 2 of his History of Greek Philosophy), Cambridge UP, 1965. Strauss, Leo; Gildin, Hilail (1989). "Progress or Return? The Contemporary Crisis in Western Education". An Introduction to Political Philosophy: Ten Essays by Leo Strauss. Wayne State University Press (published August 1, 1989). p. 209. ISBN 978-0814319024. Archived from the original on May 31, 2022. Retrieved May 30, 2022. O'Grady, Patricia F. (2016). Thales of Miletus: The Beginnings of Western Science and Philosophy. New York City, New York and London, England: Routledge. p. 245. ISBN 978-0-7546-0533-1. Archived from the original on March 31, 2021. Retrieved October 20, 2020. Burkert, Walter (June 1, 1972). Lore and Science in Ancient Pythagoreanism. Cambridge, Massachusetts: Harvard University Press. ISBN 978-0-674-53918-1. Archived from the original on January 29, 2018. Pullman, Bernard (1998). The Atom in the History of Human Thought. Oxford University Press. pp. 31–33. Bibcode:1998ahht.book.....P. ISBN 978-0-19-515040-7. Archived from the original on February 5, 2021. Retrieved October 20, 2020. Cohen, Henri; Lefebvre, Claire, eds. (2017). Handbook of Categorization in Cognitive Science (2nd ed.). Amsterdam, the Netherlands: Elsevier. p. 427. ISBN 978-0-08-101107-2. Archived from the original on February 5, 2021. Retrieved October 20, 2020. Lucretius (fl. 1st c. BCE) De rerum natura Margotta, Roberto (1968). The Story of Medicine. New York City, New York: Golden Press. Archived from the original on February 5, 2021. Retrieved November 18, 2020. Touwaide, Alain (2005). Glick, Thomas F.; Livesey, Steven; Wallis, Faith (eds.). Medieval Science, Technology, and Medicine: An Encyclopedia. New York City, New York and London, England: Routledge. p. 224. ISBN 978-0-415-96930-7. Archived from the original on February 6, 2021. Retrieved October 20, 2020. Leff, Samuel; Leff, Vera (1956). From Witchcraft to World Health. London, England: Macmillan. Archived from the original on February 5, 2021. Retrieved August 23, 2020. "Plato, Apology". p. 17. Archived from the original on January 29, 2018. Retrieved November 1, 2017. "Plato, Apology". p. 27. Archived from the original on January 29, 2018. Retrieved November 1, 2017. Aristotle. Nicomachean Ethics (H. Rackham ed.). 1139b. Archived from the original on March 17, 2012. Retrieved September 22, 2010. McClellan III, James E.; Dorn, Harold (2015). Science and Technology in World History: An Introduction. Baltimore, Maryland: Johns Hopkins University Press. pp. 99–100. ISBN 978-1-4214-1776-9. Archived from the original on February 6, 2021. Retrieved October 20, 2020. Graßhoff, Gerd (1990). The History of Ptolemy's Star Catalogue. Studies in the History of Mathematics and Physical Sciences. Vol. 14. New York, NY: Springer New York. doi:10.1007/978-1-4612-4468-4. ISBN 978-1-4612-8788-9. Archived from the original on May 30, 2022. Retrieved May 27, 2022. Hoffmann, Susanne M. (2017). Hipparchs Himmelsglobus (in German). Wiesbaden: Springer Fachmedien Wiesbaden. Bibcode:2017hihi.book.....H. doi:10.1007/978-3-658-18683-8. ISBN 978-3-658-18682-1. Archived from the original on May 30, 2022. Retrieved May 27, 2022. Edwards, C.H. Jr. (1979). The Historical Development of the Calculus. New York City, New York: Springer-Verlag. p. 75. ISBN 978-0-387-94313-8. Archived from the original on February 5, 2021. Retrieved October 20, 2020. Lawson, Russell M. (2004). Science in the Ancient World: An Encyclopedia. Santa Barbara, California: ABC-CLIO. pp. 190–91. ISBN 978-1-85109-539-1. Archived from the original on February 5, 2021. Retrieved October 20, 2020. Murphy, Trevor Morgan (2004). Pliny the Elder's Natural History: The Empire in the Encyclopedia. Oxford, England: Oxford University Press. p. 1. ISBN 978-0-19-926288-5. Archived from the original on February 6, 2021. Retrieved October 20, 2020. Doody, Aude (2010). Pliny's Encyclopedia: The Reception of the Natural History. Cambridge, England: Cambridge University Press. p. 1. ISBN 978-1-139-48453-4. Archived from the original on March 31, 2021. Retrieved October 20, 2020. Conner, Clifford D. (2005). A People's History of Science: Miners, Midwives, and "Low Mechanicks". New York: Nation Books. pp. 72–74. ISBN 1-56025-748-2. OCLC 62164511. Grant, Edward (1996). The Foundations of Modern Science in the Middle Ages: Their Religious, Institutional and Intellectual Contexts. Cambridge Studies in the History of Science. Cambridge University Press. pp. 7–17. ISBN 978-0-521-56762-6. Archived from the original on August 21, 2019. Retrieved November 9, 2018. Wildberg, Christian (May 1, 2018). Zalta, Edward N. (ed.). The Stanford Encyclopedia of Philosophy. Metaphysics Research Lab, Stanford University. Archived from the original on August 22, 2019. Retrieved May 1, 2018 – via Stanford Encyclopedia of Philosophy. Falcon, Andrea (2019). "Aristotle on Causality". In Zalta, Edward (ed.). Stanford Encyclopedia of Philosophy (Spring 2019 ed.). Metaphysics Research Lab, Stanford University. Archived from the original on October 9, 2020. Retrieved October 3, 2020. Grant, Edward (2007). "Islam and the eastward shift of Aristotelian natural philosophy". A History of Natural Philosophy: From the Ancient World to the Nineteenth Century. Cambridge University Press. pp. 62–67. ISBN 978-0-521-68957-1. Fisher, W.B. (William Bayne) (1968–1991). The Cambridge history of Iran. Cambridge: University Press. ISBN 978-0-521-20093-6. OCLC 745412. "Bayt al-Hikmah". Encyclopædia Britannica. Archived from the original on November 4, 2016. Retrieved November 3, 2016. Hossein Nasr, Seyyed; Leaman, Oliver, eds. (2001). History of Islamic Philosophy. Routledge. pp. 165–167. ISBN 978-0415259347. Smith, A. Mark (2001). Alhacen's Theory of Visual Perception: A Critical Edition, with English Translation and Commentary, of the First Three Books of Alhacen's De Aspectibus, the Medieval Latin Version of Ibn al-Haytham's Kitāb al-Manāẓir, 2 vols. Transactions of the American Philosophical Society. Vol. 91. Philadelphia: American Philosophical Society. ISBN 978-0-87169-914-5. OCLC 47168716. Toomer, G.J. (1964). "Reviewed work: Ibn al-Haythams Weg zur Physik, Matthias Schramm". Isis. 55 (4): 463–65. doi:10.1086/349914. JSTOR 228328. See p. 464: "Schramm sums up [Ibn Al-Haytham's] achievement in the development of scientific method.", p. 465: "Schramm has demonstrated .. beyond any dispute that Ibn al-Haytham is a major figure in the Islamic scientific tradition, particularly in the creation of experimental techniques." p. 465: "only when the influence of Ibn al-Haytham and others on the mainstream of later medieval physical writings has been seriously investigated can Schramm's claim that Ibn al-Haytham was the true founder of modern physics be evaluated." Cohen, H. Floris (2010). "Greek nature knowledge transplanted: The Islamic world". How modern science came into the world. Four civilizations, one 17th-century breakthrough (2nd ed.). Amsterdam: Amsterdam University Press. pp. 99–156. ISBN 978-90-8964-239-4. Selin, Helaine, ed. (2006). Encyclopaedia of the History of Science, Technology, and Medicine in Non-Western Cultures. Springer. pp. 155–156. Bibcode:2008ehst.book.....S. ISBN 978-1-4020-4559-2. Russell, Josiah C. (1959). "Gratian, Irnerius, and the Early Schools of Bologna". The Mississippi Quarterly. 12 (4): 168–188. JSTOR 26473232. Archived from the original on May 27, 2022. Retrieved May 27, 2022 – via JSTOR. Perhaps even as early as 1088 (the date officially set for the founding of the University) "St. Albertus Magnus | German theologian, scientist, and philosopher". Archived from the original on October 28, 2017. Retrieved October 27, 2017. Numbers, Ronald (2009). Galileo Goes to Jail and Other Myths about Science and Religion. Harvard University Press. p. 45. ISBN 978-0-674-03327-6. Archived from the original on January 20, 2021. Retrieved March 27, 2018. Smith, A. Mark (1981). "Getting the Big Picture in Perspectivist Optics". Isis. 72 (4): 568–89. doi:10.1086/352843. JSTOR 231249. PMID 7040292. S2CID 27806323. Goldstein, Bernard R (2016). "Copernicus and the Origin of his Heliocentric System" (PDF). Journal for the History of Astronomy. 33 (3): 219–35. doi:10.1177/002182860203300301. S2CID 118351058. Archived from the original (PDF) on April 12, 2020. Retrieved April 12, 2020. Cohen, H. Floris (2010). "Greek nature knowledge transplanted and more: Renaissance Europe". How modern science came into the world. Four civilizations, one 17th-century breakthrough (2nd ed.). Amsterdam: Amsterdam University Press. pp. 99–156. ISBN 978-90-8964-239-4. Koestler, Arthur (1990) [1959]. The Sleepwalkers: A History of Man's Changing Vision of the Universe. London: Penguin Books. p. 1. ISBN 0-14-019246-8. van Helden, Al (1995). "Pope Urban VIII". The Galileo Project. Archived from the original on November 11, 2016. Retrieved November 3, 2016. Gingerich, Owen (1975). "Copernicus and the Impact of Printing". Vistas in Astronomy. 17 (1): 201–218. Bibcode:1975VA.....17..201G. doi:10.1016/0083-6656(75)90061-6. Zagorin, Perez (1998). Francis Bacon. Princeton: Princeton University Press. p. 84. ISBN 978-0-691-00966-7. Davis, Philip J.; Hersh, Reuben (1986). Descartes' Dream: The World According to Mathematics. Cambridge, MA: Harcourt Brace Jovanovich. Gribbin, John (2002). Science: A History 1543–2001. Allen Lane. p. 241. ISBN 978-0-7139-9503-9. Although it was just one of the many factors in the Enlightenment, the success of Newtonian physics in providing a mathematical description of an ordered world clearly played a big part in the flowering of this movement in the eighteenth century "Gottfried Leibniz – Biography". Maths History. Archived from the original on July 11, 2017. Retrieved March 2, 2021. Freudenthal, Gideon; McLaughlin, Peter (May 20, 2009). The Social and Economic Roots of the Scientific Revolution: Texts by Boris Hessen and Henryk Grossmann. Springer Science & Business Media. ISBN 978-1-4020-9604-4. Archived from the original on January 19, 2020. Retrieved July 25, 2018. Goddard Bergin, Thomas; Speake, Jennifer, eds. (1987). Encyclopedia of the Renaissance. Facts on File (published December 1, 1987). ISBN 978-0816013159. van Horn Melton, James (2001). The Rise of the Public in Enlightenment Europe. Cambridge University Press. pp. 82–83. doi:10.1017/CBO9780511819421. ISBN 978-0511819421. Archived from the original on January 20, 2022. Retrieved May 27, 2022. "The Scientific Revolution and the Enlightenment (1500–1780)" (PDF). Retrieved January 29, 2024. "Scientific Revolution | Definition, History, Scientists, Inventions, & Facts". Britannica. Retrieved January 29, 2024. Madigan M, Martinko J, eds. (2006). Brock Biology of Microorganisms (11th ed.). Prentice Hall. ISBN 978-0131443297. Guicciardini, N. (1999). Reading the Principia: The Debate on Newton's Methods for Natural Philosophy from 1687 to 1736. New York: Cambridge University Press. ISBN 978-0521640664. Calisher, CH (2007). "Taxonomy: what's in a name? Doesn't a rose by any other name smell as sweet?". Croatian Medical Journal. 48 (2): 268–270. PMC 2080517. PMID 17436393. Darrigol, Olivier (2000). Electrodynamics from Ampère to Einstein. New York: Oxford University Press. ISBN 0198505949. Olby, R.C.; Cantor, G.N.; Christie, J.R.R.; Hodge, M.J.S. (1990). Companion to the History of Modern Science. London: Routledge. p. 265. Magnusson, Magnus (November 10, 2003). "Review of James Buchan, Capital of the Mind: how Edinburgh Changed the World". New Statesman. Archived from the original on June 6, 2011. Retrieved April 27, 2014. Swingewood, Alan (1970). "Origins of Sociology: The Case of the Scottish Enlightenment". The British Journal of Sociology. 21 (2): 164–180. doi:10.2307/588406. JSTOR 588406. Fry, Michael (1992). Adam Smith's Legacy: His Place in the Development of Modern Economics. Paul Samuelson, Lawrence Klein, Franco Modigliani, James M. Buchanan, Maurice Allais, Theodore Schultz, Richard Stone, James Tobin, Wassily Leontief, Jan Tinbergen. Routledge. ISBN 978-0-415-06164-3. Lightman, Bernard (2011). "13. Science and the Public". In Shank, Michael; Numbers, Ronald; Harrison, Peter (eds.). Wrestling with Nature: From Omens to Science. Chicago: University of Chicago Press. p. 367. ISBN 978-0-226-31783-0. Leahey, Thomas Hardy (2018). "The psychology of consciousness". A History of Psychology: From Antiquity to Modernity (8th ed.). New York, NY: Routledge. pp. 219–253. ISBN 978-1-138-65242-2. Padian, Kevin (2008). "Darwin's enduring legacy". Nature. 451 (7179): 632–634. Bibcode:2008Natur.451..632P. doi:10.1038/451632a. PMID 18256649. Henig, Robin Marantz (2000). The monk in the garden: the lost and found genius of Gregor Mendel, the father of genetics. pp. 134–138. Miko, Ilona (2008). "Gregor Mendel's principles of inheritance form the cornerstone of modern genetics. So just what are they?". Nature Education. 1 (1): 134. Archived from the original on July 19, 2019. Retrieved May 9, 2021. Rocke, Alan J. (2005). "In Search of El Dorado: John Dalton and the Origins of the Atomic Theory". Social Research. 72 (1): 125–158. doi:10.1353/sor.2005.0003. JSTOR 40972005. S2CID 141350239. Reichl, Linda (1980). A Modern Course in Statistical Physics. Edward Arnold. ISBN 0-7131-2789-9. Rao, Y. V. C. (1997). Chemical Engineering Thermodynamics. Universities Press. p. 158. ISBN 978-81-7371-048-3. Heidrich, M. (2016). "Bounded energy exchange as an alternative to the third law of thermodynamics". Annals of Physics. 373: 665–681. Bibcode:2016AnPhy.373..665H. doi:10.1016/j.aop.2016.07.031. Archived from the original on January 15, 2019. Retrieved May 29, 2022. Mould, Richard F. (1995). A century of X-rays and radioactivity in medicine: with emphasis on photographic records of the early years (Reprint. with minor corr ed.). Bristol: Inst. of Physics Publ. p. 12. ISBN 978-0-7503-0224-1. Estreicher, Tadeusz (1938). "Curie, Maria ze Skłodowskich". Polski słownik biograficzny, vol. 4 (in Polish). p. 113. Thomson, J.J. (1897). "Cathode Rays". Philosophical Magazine. 44 (269): 293–316. doi:10.1080/14786449708621070. Retrieved February 24, 2022. Goyotte, Dolores (2017). "The Surgical Legacy of World War II. Part II: The age of antibiotics" (PDF). The Surgical Technologist. 109: 257–264. Archived (PDF) from the original on May 5, 2021. Retrieved January 8, 2021. Erisman, Jan Willem; MA Sutton; J Galloway; Z Klimont; W Winiwarter (October 2008). "How a century of ammonia synthesis changed the world". Nature Geoscience. 1 (10): 636–639. Bibcode:2008NatGe...1..636E. doi:10.1038/ngeo325. S2CID 94880859. Archived from the original on July 23, 2010. Retrieved October 22, 2010. Emmett, Robert; Zelko, Frank (2014). Emmett, Rob; Zelko, Frank (eds.). "Minding the Gap: Working Across Disciplines in Environmental Studies". Environment & Society Portal. RCC Perspectives no. 2. doi:10.5282/rcc/6313. Archived from the original on January 21, 2022. Furner, Jonathan (June 1, 2003). "Little Book, Big Book: Before and After Little Science, Big Science: A Review Article, Part I". Journal of Librarianship and Information Science. 35 (2): 115–125. doi:10.1177/0961000603352006. S2CID 34844169. Kraft, Chris; James Schefter (2001). Flight: My Life in Mission Control. New York: Dutton. pp. 3–5. ISBN 0-525-94571-7. Kahn, Herman (1962). Thinking about the Unthinkable. Horizon Press. Shrum, Wesley (2007). Structures of scientific collaboration. Joel Genuth, Ivan Chompalov. Cambridge, Mass.: MIT Press. ISBN 978-0-262-28358-8. OCLC 166143348. Archived from the original on July 30, 2022. Retrieved May 31, 2022. Rosser, Sue V. (March 12, 2012). Breaking into the Lab: Engineering Progress for Women in Science. New York: New York University Press. p. 7. ISBN 978-0-8147-7645-2. Penzias, A. A. (2006). "The origin of elements" (PDF). Science. 205 (4406). Nobel Foundation: 549–54. doi:10.1126/science.205.4406.549. PMID 17729659. Archived (PDF) from the original on January 17, 2011. Retrieved October 4, 2006. Weinberg, S. (1972). Gravitation and Cosmology. John Whitney & Sons. pp. 495–464. ISBN 978-0-471-92567-5. Futuyma, Douglas J.; Kirkpatrick, Mark (2017). "Chapter 1: Evolutionary Biology". Evolution (4th ed.). Sinauer. pp. 3–26. ISBN 978-1605356051. Archived from the original on May 31, 2022. Retrieved May 30, 2022. Miller, Arthur I. (1981). Albert Einstein's special theory of relativity. Emergence (1905) and early interpretation (1905–1911). Reading: Addison–Wesley. ISBN 978-0-201-04679-3. ter Haar, D. (1967). The Old Quantum Theory. Pergamon Press. pp. 206. ISBN 978-0-08-012101-7. von Bertalanffy, Ludwig (1972). "The History and Status of General Systems Theory". The Academy of Management Journal. 15 (4): 407–26. JSTOR 255139. Naidoo, Nasheen; Pawitan, Yudi; Soong, Richie; Cooper, David N.; Ku, Chee-Seng (October 2011). "Human genetics and genomics a decade after the release of the draft sequence of the human genome". Human Genomics. 5 (6): 577–622. doi:10.1186/1479-7364-5-6-577. PMC 3525251. PMID 22155605. Rashid, S. Tamir; Alexander, Graeme J.M. (March 2013). "Induced pluripotent stem cells: from Nobel Prizes to clinical applications". Journal of Hepatology. 58 (3): 625–629. doi:10.1016/j.jhep.2012.10.026. ISSN 1600-0641. PMID 23131523. O'Luanaigh, C. (March 14, 2013). "New results indicate that new particle is a Higgs boson" (Press release). CERN. Archived from the original on October 20, 2015. Retrieved October 9, 2013. Abbott, B.P.; Abbott, R.; Abbott, T.D.; Acernese, F.; Ackley, K.; Adams, C.; Adams, T.; Addesso, P.; Adhikari, R.X.; Adya, V.B.; Affeldt, C.; Afrough, M.; Agarwal, B.; Agathos, M.; Agatsuma, K.; Aggarwal, N.; Aguiar, O.D.; Aiello, L.; Ain, A.; Ajith, P.; Allen, B.; Allen, G.; Allocca, A.; Altin, P.A.; Amato, A.; Ananyeva, A.; Anderson, S.B.; Anderson, W.G.; Angelova, S.V.; et al. (2017). "Multi-messenger Observations of a Binary Neutron Star Merger". The Astrophysical Journal. 848 (2): L12. arXiv:1710.05833. Bibcode:2017ApJ...848L..12A. doi:10.3847/2041-8213/aa91c9. S2CID 217162243. Cho, Adrian (2017). "Merging neutron stars generate gravitational waves and a celestial light show". Science. doi:10.1126/science.aar2149. "Media Advisory: First Results from the Event Horizon Telescope to be Presented on April 10th | Event Horizon Telescope". April 20, 2019. Archived from the original on April 20, 2019. Retrieved September 21, 2021. "Scientific Method: Relationships Among Scientific Paradigms". Seed Magazine. March 7, 2007. Archived from the original on November 1, 2016. Retrieved November 4, 2016. Bunge, Mario Augusto (1998). Philosophy of Science: From Problem to Theory. Transaction Publishers. p. 24. ISBN 978-0-7658-0413-6. Popper, Karl R. (2002a) [1959]. "A survey of some fundamental problems". The Logic of Scientific Discovery. New York: Routledge Classics. pp. 3–26. ISBN 978-0-415-27844-7. OCLC 59377149. Gauch, Hugh G. Jr. (2003). "Science in perspective". Scientific Method in Practice. Cambridge, United Kingdom: Cambridge University Press. pp. 21–73. ISBN 978-0-521-01708-4. Archived from the original on December 25, 2020. Retrieved September 3, 2018. Oglivie, Brian W. (2008). "Introduction". The Science of Describing: Natural History in Renaissance Europe (Paperback ed.). Chicago: University of Chicago Press. pp. 1–24. ISBN 978-0-226-62088-6. "Natural History". Princeton University WordNet. Archived from the original on March 3, 2012. Retrieved October 21, 2012. "Formal Sciences: Washington and Lee University". Washington and Lee University. Archived from the original on May 14, 2021. Retrieved May 14, 2021. A "formal science" is an area of study that uses formal systems to generate knowledge such as in Mathematics and Computer Science. Formal sciences are important subjects because all of quantitative science depends on them. "formal system". Encyclopædia Britannica. Archived from the original on April 29, 2008. Retrieved May 30, 2022. Tomalin, Marcus (2006). Linguistics and the Formal Sciences. Löwe, Benedikt (2002). "The Formal Sciences: Their Scope, Their Foundations, and Their Unity". Synthese. 133: 5–11. doi:10.1023/a:1020887832028. S2CID 9272212. Bill, Thompson (2007). "2.4 Formal Science and Applied Mathematics". The Nature of Statistical Evidence. Lecture Notes in Statistics. Vol. 189. Springer. p. 15. Bunge, Mario (1998). "The Scientific Approach". Philosophy of Science: Volume 1, From Problem to Theory. Vol. 1 (revised ed.). New York: Routledge. pp. 3–50. ISBN 978-0-7658-0413-6. Mujumdar, Anshu Gupta; Singh, Tejinder (2016). "Cognitive science and the connection between physics and mathematics". In Aguirre, Anthony; Foster, Brendan (eds.). Trick or Truth?: The Mysterious Connection Between Physics and Mathematics. The Frontiers Collection. Switzerland: SpringerNature. pp. 201–218. ISBN 978-3-319-27494-2. "About the Journal". Journal of Mathematical Physics. Archived from the original on October 3, 2006. Retrieved October 3, 2006. Restrepo, G. (2016). "Mathematical chemistry, a new discipline". In Scerri, E.; Fisher, G. (eds.). Essays in the philosophy of chemistry. New York, UK: Oxford University Press. pp. 332–351. ISBN 978-0-19-049459-9. Archived from the original on June 10, 2021. "What is mathematical biology". Centre for Mathematical Biology, University of Bath. Archived from the original on September 23, 2018. Retrieved June 7, 2018. Johnson, Tim (September 1, 2009). "What is financial mathematics?". +Plus Magazine. Archived from the original on April 8, 2022. Retrieved March 1, 2021. Varian, Hal (1997). "What Use Is Economic Theory?". In D'Autume, A.; Cartelier, J. (eds.). Is Economics Becoming a Hard Science?. Edward Elgar. Pre-publication. Archived June 25, 2006, at the Wayback Machine. Retrieved April 1, 2008. Abraham, Reem Rachel (2004). "Clinically oriented physiology teaching: strategy for developing critical-thinking skills in undergraduate medical students". Advances in Physiology Education. 28 (3): 102–04. doi:10.1152/advan.00001.2004. PMID 15319191. S2CID 21610124. "Cambridge Dictionary". Cambridge University Press. Archived from the original on August 19, 2019. Retrieved March 25, 2021. Brooks, Harvey (September 1, 1994). "The relationship between science and technology" (PDF). Research Policy. Special Issue in Honor of Nathan Rosenberg. 23 (5): 477–486. doi:10.1016/0048-7333(94)01001-3. ISSN 0048-7333. Firth, John (2020). "Science in medicine: when, how, and what". Oxford textbook of medicine. Oxford: Oxford University Press. ISBN 978-0-19-874669-0. Saunders, J. (June 2000). "The practice of clinical medicine as an art and as a science". Med Humanit. 26 (1): 18–22. doi:10.1136/mh.26.1.18. PMC 1071282. PMID 12484313. S2CID 73306806. Davis, Bernard D. (March 2000). "Limited scope of science". Microbiology and Molecular Biology Reviews. 64 (1): 1–12. doi:10.1128/MMBR.64.1.1-12.2000. PMC 98983. PMID 10704471 & "Technology" in Davis, Bernard (March 2000). "The scientist's world". Microbiology and Molecular Biology Reviews. 64 (1): 1–12. doi:10.1128/MMBR.64.1.1-12.2000. PMC 98983. PMID 10704471. McCormick, James (2001). "Scientific medicine—fact of fiction? The contribution of science to medicine". Occasional Paper (Royal College of General Practitioners) (80): 3–6. PMC 2560978. PMID 19790950. Breznau, Nate (2022). "Integrating Computer Prediction Methods in Social Science: A Comment on Hofman et al. (2021)". Social Science Computer Review. 40 (3): 844–853. doi:10.1177/08944393211049776. S2CID 248334446. Hofman, Jake M.; Watts, Duncan J.; Athey, Susan; Garip, Filiz; Griffiths, Thomas L.; Kleinberg, Jon; Margetts, Helen; Mullainathan, Sendhil; Salganik, Matthew J.; Vazire, Simine; Vespignani, Alessandro (July 2021). "Integrating explanation and prediction in computational social science". Nature. 595 (7866): 181–188. Bibcode:2021Natur.595..181H. doi:10.1038/s41586-021-03659-0. ISSN 1476-4687. PMID 34194044. S2CID 235697917. Archived from the original on September 25, 2021. Retrieved September 25, 2021. Nissani, M. (1995). "Fruits, Salads, and Smoothies: A Working definition of Interdisciplinarity". The Journal of Educational Thought. 29 (2): 121–128. JSTOR 23767672. Moody G (2004). Digital Code of Life: How Bioinformatics is Revolutionizing Science, Medicine, and Business. John Wiley & Sons. p. vii. ISBN 978-0-471-32788-2. Ausburg, Tanya (2006). Becoming Interdisciplinary: An Introduction to Interdisciplinary Studies (2nd ed.). New York: Kendall/Hunt Publishing. Dawkins, Richard (May 10, 2006). "To Live at All Is Miracle Enough". RichardDawkins.net. Archived from the original on January 19, 2012. Retrieved February 5, 2012. di Francia, Giuliano Toraldo (1976). "The method of physics". The Investigation of the Physical World. Cambridge, United Kingdom: Cambridge University Press. pp. 1–52. ISBN 978-0-521-29925-1. The amazing point is that for the first time since the discovery of mathematics, a method has been introduced, the results of which have an intersubjective value! Popper, Karl R. (2002e) [1959]. "The problem of the empirical basis". The Logic of Scientific Discovery. New York: Routledge Classics. pp. 3–26. ISBN 978-0-415-27844-7. OCLC 59377149. Diggle, Peter J.; Chetwynd, Amanda G. (2011). Statistics and Scientific Method: An Introduction for Students and Researchers. Oxford University Press. pp. 1, 2. ISBN 978-0199543182. Wilson, Edward (1999). Consilience: The Unity of Knowledge. New York: Vintage. ISBN 978-0-679-76867-8. Fara, Patricia (2009). "Decisions". Science: A Four Thousand Year History. Oxford, United Kingdom: Oxford University Press. p. 408. ISBN 978-0-19-922689-4. Aldrich, John (1995). "Correlations Genuine and Spurious in Pearson and Yule". Statistical Science. 10 (4): 364–376. doi:10.1214/ss/1177009870. JSTOR 2246135. Nola, Robert; Irzik, Gürol (2005k). "naive inductivism as a methodology in science". Philosophy, science, education and culture. Science & technology education library. Vol. 28. Springer. pp. 207–230. ISBN 978-1-4020-3769-6. Nola, Robert; Irzik, Gürol (2005j). "The aims of science and critical inquiry". Philosophy, science, education and culture. Science & technology education library. Vol. 28. Springer. pp. 207–230. ISBN 978-1-4020-3769-6. van Gelder, Tim (1999). ""Heads I win, tails you lose": A Foray Into the Psychology of Philosophy" (PDF). University of Melbourne. Archived from the original (PDF) on April 9, 2008. Retrieved March 28, 2008. Pease, Craig (September 6, 2006). "Chapter 23. Deliberate bias: Conflict creates bad science". Science for Business, Law and Journalism. Vermont Law School. Archived from the original on June 19, 2010. Shatz, David (2004). Peer Review: A Critical Inquiry. Rowman & Littlefield. ISBN 978-0-7425-1434-8. OCLC 54989960. Krimsky, Sheldon (2003). Science in the Private Interest: Has the Lure of Profits Corrupted the Virtue of Biomedical Research. Rowman & Littlefield. ISBN 978-0-7425-1479-9. OCLC 185926306. Bulger, Ruth Ellen; Heitman, Elizabeth; Reiser, Stanley Joel (2002). The Ethical Dimensions of the Biological and Health Sciences (2nd ed.). Cambridge University Press. ISBN 978-0-521-00886-0. OCLC 47791316. Backer, Patricia Ryaby (October 29, 2004). "What is the scientific method?". San Jose State University. Archived from the original on April 8, 2008. Retrieved March 28, 2008. Ziman, John (1978c). "Common observation". Reliable knowledge: An exploration of the grounds for belief in science. Cambridge: Cambridge University Press. pp. 42–76. ISBN 978-0-521-22087-3. Ziman, J.M. (1980). "The proliferation of scientific literature: a natural process". Science. 208 (4442): 369–71. Bibcode:1980Sci...208..369Z. doi:10.1126/science.7367863. PMID 7367863. Subramanyam, Krishna; Subramanyam, Bhadriraju (1981). Scientific and Technical Information Resources. CRC Press. ISBN 978-0-8247-8297-9. OCLC 232950234. Bush, Vannevar (July 1945). "Science the Endless Frontier". National Science Foundation. Archived from the original on November 7, 2016. Retrieved November 4, 2016. Schooler, J. W. (2014). "Metascience could rescue the 'replication crisis'". Nature. 515 (7525): 9. Bibcode:2014Natur.515....9S. doi:10.1038/515009a. PMID 25373639. Pashler, Harold; Wagenmakers, Eric Jan (2012). "Editors' Introduction to the Special Section on Replicability in Psychological Science: A Crisis of Confidence?". Perspectives on Psychological Science. 7 (6): 528–530. doi:10.1177/1745691612465253. PMID 26168108. S2CID 26361121. Ioannidis, John P. A.; Fanelli, Daniele; Dunne, Debbie Drake; Goodman, Steven N. (October 2, 2015). "Meta-research: Evaluation and Improvement of Research Methods and Practices". PLOS Biology. 13 (10): –1002264. doi:10.1371/journal.pbio.1002264. ISSN 1545-7885. PMC 4592065. PMID 26431313. Hansson, Sven Ove; Zalta, Edward N. (September 3, 2008). "Science and Pseudoscience". Stanford Encyclopedia of Philosophy. Section 2: The "science" of pseudoscience. Archived from the original on October 29, 2021. Retrieved May 28, 2022. Shermer M (1997). Why people believe weird things: pseudoscience, superstition, and other confusions of our time. New York: W. H. Freeman and Company. p. 17. ISBN 978-0-7167-3090-3. Feynman, Richard (1974). "Cargo Cult Science". Center for Theoretical Neuroscience. Columbia University. Archived from the original on March 4, 2005. Retrieved November 4, 2016. Novella, Steven (2018). The Skeptics' Guide to the Universe: How to Know What's Really Real in a World Increasingly Full of Fake. Hodder & Stoughton. p. 162. ISBN 978-1473696419. "Coping with fraud" (PDF). The COPE Report 1999: 11–18. Archived from the original (PDF) on September 28, 2007. Retrieved July 21, 2011. It is 10 years, to the month, since Stephen Lock ... Reproduced with kind permission of the Editor, The Lancet. Godfrey-Smith, Peter (2003c). "Induction and confirmation". Theory and Reality: An Introduction to the Philosophy of Science. Chicago: University of Chicago. pp. 39–56. ISBN 978-0-226-30062-7. Godfrey-Smith, Peter (2003o). "Empiricism, naturalism, and scientific realism?". Theory and Reality: An Introduction to the Philosophy of Science. Chicago: University of Chicago. pp. 219–232. ISBN 978-0-226-30062-7. Godfrey-Smith, Peter (2003b). "Logic plus empiricism". Theory and Reality: An Introduction to the Philosophy of Science. Chicago: University of Chicago. pp. 19–38. ISBN 978-0-226-30062-7. Godfrey-Smith, Peter (2003d). "Popper: Conjecture and refutation". Theory and Reality: An Introduction to the Philosophy of Science. Chicago: University of Chicago. pp. 57–74. ISBN 978-0-226-30062-7. Godfrey-Smith, Peter (2003g). "Lakatos, Laudan, Feyerabend, and frameworks". Theory and Reality: An Introduction to the Philosophy of Science. Chicago: University of Chicago. pp. 102–121. ISBN 978-0-226-30062-7. Popper, Karl (1972). Objective Knowledge. Newton-Smith, W.H. (1994). The Rationality of Science. London: Routledge. p. 30. ISBN 978-0-7100-0913-5. Votsis, I. (2004). The Epistemological Status of Scientific Theories: An Investigation of the Structural Realist Account (PhD Thesis). University of London, London School of Economics. p. 39. Bird, Alexander (2013). Zalta, Edward N. (ed.). "Thomas Kuhn". Stanford Encyclopedia of Philosophy. Archived from the original on July 15, 2020. Retrieved October 26, 2015. Kuhn, Thomas S. (1970). The Structure of Scientific Revolutions (2nd ed.). University of Chicago Press. p. 206. ISBN 978-0-226-45804-5. Archived from the original on October 19, 2021. Retrieved May 30, 2022. Godfrey-Smith, Peter (2003). "Naturalistic philosophy in theory and practice". Theory and Reality: An Introduction to the Philosophy of Science. Chicago: University of Chicago. pp. 149–162. ISBN 978-0-226-30062-7. Brugger, E. Christian (2004). "Casebeer, William D. Natural Ethical Facts: Evolution, Connectionism, and Moral Cognition". The Review of Metaphysics. 58 (2). Kornfeld, W; Hewitt, CE (1981). "The Scientific Community Metaphor" (PDF). IEEE Transactions on Systems, Man, and Cybernetics. 11 (1): 24–33. doi:10.1109/TSMC.1981.4308575. hdl:1721.1/5693. S2CID 1322857. Archived (PDF) from the original on April 8, 2016. Retrieved May 26, 2022. "Eusocial climbers" (PDF). E.O. Wilson Foundation. Archived (PDF) from the original on April 27, 2019. Retrieved September 3, 2018. But he's not a scientist, he's never done scientific research. My definition of a scientist is that you can complete the following sentence: 'he or she has shown that...'," Wilson says. "Our definition of a scientist". Science Council. Archived from the original on August 23, 2019. Retrieved September 7, 2018. A scientist is someone who systematically gathers and uses research and evidence, making a hypothesis and testing it, to gain and share understanding and knowledge. Cyranoski, David; Gilbert, Natasha; Ledford, Heidi; Nayar, Anjali; Yahia, Mohammed (2011). "Education: The PhD factory". Nature. 472 (7343): 276–79. Bibcode:2011Natur.472..276C. doi:10.1038/472276a. PMID 21512548. Kwok, Roberta (2017). "Flexible working: Science in the gig economy". Nature. 550: 419–21. doi:10.1038/nj7677-549a. Woolston, Chris (2007). Editorial (ed.). "Many junior scientists need to take a hard look at their job prospects". Nature. 550: 549–552. doi:10.1038/nj7677-549a. Lee, Adrian; Dennis, Carina; Campbell, Phillip (2007). "Graduate survey: A love–hurt relationship". Nature. 550 (7677): 549–52. doi:10.1038/nj7677-549a. Cyranoski, David; Gilbert, Natasha; Ledford, Heidi; Nayar, Anjali; Yahia, Mohammed (2011). "Education: The PhD factory". Nature. 472 (7343): 276–279. Bibcode:2011Natur.472..276C. doi:10.1038/472276a. PMID 21512548. Kwok, Roberta (2017). "Flexible working: Science in the gig economy". Nature. 550: 419–421. doi:10.1038/nj7677-549a. Lee, Adrian; Dennis, Carina; Campbell, Phillip (2007). "Graduate survey: A love–hurt relationship". Nature. 550 (7677): 549–552. doi:10.1038/nj7677-549a. Whaley, Leigh Ann (2003). Women's History as Scientists. Santa Barbara, California: ABC-CLIO, INC. Spanier, Bonnie (1995). "From Molecules to Brains, Normal Science Supports Sexist Beliefs about Difference". Im/partial Science: Gender Identity in Molecular Biology. Indiana University Press. ISBN 978-0-253-20968-9. Parrott, Jim (August 9, 2007). "Chronicle for Societies Founded from 1323 to 1599". Scholarly Societies Project. Archived from the original on January 6, 2014. Retrieved September 11, 2007. "The Environmental Studies Association of Canada – What is a Learned Society?". Archived from the original on May 29, 2013. Retrieved May 10, 2013. "Learned societies & academies". Archived from the original on June 3, 2014. Retrieved May 10, 2013. "Learned Societies, the key to realising an open access future?". Impact of Social Sciences. London School of Economics. June 24, 2019. Retrieved January 22, 2023. "Accademia Nazionale dei Lincei" (in Italian). 2006. Archived from the original on February 28, 2010. Retrieved September 11, 2007. "Prince of Wales opens Royal Society's refurbished building". The Royal Society. July 7, 2004. Archived from the original on April 9, 2015. Retrieved December 7, 2009. Meynell, G.G. "The French Academy of Sciences, 1666–91: A reassessment of the French Académie royale des sciences under Colbert (1666–83) and Louvois (1683–91)". Archived from the original on January 18, 2012. Retrieved October 13, 2011. ITS. "Founding of the National Academy of Sciences". .nationalacademies.org. Archived from the original on February 3, 2013. Retrieved March 12, 2012. "The founding of the Kaiser Wilhelm Society (1911)". Max-Planck-Gesellschaft. Archived from the original on March 2, 2022. Retrieved May 30, 2022. "Introduction". Chinese Academy of Sciences. Archived from the original on March 31, 2022. Retrieved May 31, 2022. "Two main Science Councils merge to address complex global challenges". UNESCO. July 5, 2018. Archived from the original on July 12, 2021. Retrieved October 21, 2018. Stockton, Nick (October 7, 2014). "How did the Nobel Prize become the biggest award on Earth?". Wired. Archived from the original on June 19, 2019. Retrieved September 3, 2018. "Main Science and Technology Indicators – 2008-1" (PDF). OECD. Archived from the original (PDF) on February 15, 2010. OECD Science, Technology and Industry Scoreboard 2015: Innovation for growth and society. OECD. 2015. p. 156. doi:10.1787/sti_scoreboard-2015-en. ISBN 978-9264239784. Archived from the original on May 25, 2022. Retrieved May 28, 2022 – via oecd-ilibrary.org. Kevles, Daniel (1977). "The National Science Foundation and the Debate over Postwar Research Policy, 1942-1945". Isis. 68 (241): 4–26. doi:10.1086/351711. PMID 320157. S2CID 32956693. "Argentina, National Scientific and Technological Research Council (CONICET)". International Science Council. Archived from the original on May 16, 2022. Retrieved May 31, 2022. Innis, Michelle (May 17, 2016). "Australia to Lay Off Leading Scientist on Sea Levels". The New York Times. ISSN 0362-4331. Archived from the original on May 7, 2021. Retrieved May 31, 2022. "Le CNRS recherche 10.000 passionnés du blob". Le Figaro (in French). October 20, 2021. Archived from the original on April 27, 2022. Retrieved May 31, 2022. Bredow, Rafaela von (December 18, 2021). "How a Prestigious Scientific Organization Came Under Suspicion of Treating Women Unequally". Der Spiegel. ISSN 2195-1349. Archived from the original on May 29, 2022. Retrieved May 31, 2022. "En espera de una "revolucionaria" noticia sobre Sagitario A*, el agujero negro supermasivo en el corazón de nuestra galaxia". ELMUNDO (in Spanish). May 12, 2022. Archived from the original on May 13, 2022. Retrieved May 31, 2022. Fletcher, Anthony C.; Bourne, Philip E. (September 27, 2012). "Ten Simple Rules To Commercialize Scientific Research". PLOS Computational Biology. 8 (9): e1002712. Bibcode:2012PLSCB...8E2712F. doi:10.1371/journal.pcbi.1002712. ISSN 1553-734X. PMC 3459878. PMID 23028299. Marburger, John Harmen III (February 10, 2015). Science policy up close. Crease, Robert P. Cambridge, Massachusetts: Harvard University Press. ISBN 978-0-674-41709-0. OCLC 875999943. Benneworth, Paul; Jongbloed, Ben W. (July 31, 2009). "Who matters to universities? A stakeholder perspective on humanities, arts and social sciences valorisation" (PDF). Higher Education. 59 (5): 567–588. doi:10.1007/s10734-009-9265-2. ISSN 0018-1560. Dickson, David (October 11, 2004). "Science journalism must keep a critical edge". Science and Development Network. Archived from the original on June 21, 2010. Mooney, Chris (November–December 2004). "Blinded By Science, How 'Balanced' Coverage Lets the Scientific Fringe Hijack Reality". Columbia Journalism Review. Vol. 43, no. 4. Archived from the original on January 17, 2010. Retrieved February 20, 2008. McIlwaine, S.; Nguyen, D.A. (2005). "Are Journalism Students Equipped to Write About Science?". Australian Studies in Journalism. 14: 41–60. Archived from the original on August 1, 2008. Retrieved February 20, 2008. Webb, Sarah (December 2013). "Popular science: Get the word out". Nature. 504 (7478): 177–9. doi:10.1038/nj7478-177a. PMID 24312943. Wilde, Fran (January 21, 2016). "How Do You Like Your Science Fiction? Ten Authors Weigh In On 'Hard' vs. 'Soft' SF". Tor.com. Archived from the original on April 4, 2019. Retrieved April 4, 2019. Petrucci, Mario. "Creative Writing – Science". Archived from the original on January 6, 2009. Retrieved April 27, 2008. Tyson, Alec; Funk, Cary; Kennedy, Brian; Johnson, Courtney (September 15, 2021). "Majority in U.S. Says Public Health Benefits of COVID-19 Restrictions Worth the Costs, Even as Large Shares Also See Downsides". Pew Research Center Science & Society. Retrieved August 4, 2022. Kennedy, Brian. "U.S. concern about climate change is rising, but mainly among Democrats". Pew Research Center. Retrieved August 4, 2022. Philipp-Muller, Aviva; Lee, Spike W. S.; Petty, Richard E. (July 26, 2022). "Why are people antiscience, and what can we do about it?". Proceedings of the National Academy of Sciences. 119 (30): e2120755119. Bibcode:2022PNAS..11920755P. doi:10.1073/pnas.2120755119. ISSN 0027-8424. PMC 9335320. PMID 35858405. Gauchat, Gordon William (2008). "A Test of Three Theories of Anti-Science Attitudes". Sociological Focus. 41 (4): 337–357. doi:10.1080/00380237.2008.10571338. S2CID 144645723. Poushter, Jacob; Fagan, Moira; Gubbala, Sneha (August 31, 2022). "Climate Change Remains Top Global Threat Across 19-Country Survey". Pew Research Center's Global Attitudes Project. Retrieved September 5, 2022. McRaney, David (2022). How Minds Change: The Surprising Science of Belief, Opinion, and Persuasion. [New York, NY]: Portfolio/Penguin. ISBN 978-0-593-19029-6. OCLC 1322437138. McGreal, Chris (October 26, 2021). "Revealed: 60% of Americans say oil firms are to blame for the climate crisis". The Guardian. Archived from the original on October 26, 2021. Source: Guardian/Vice/CCN/YouGov poll. Note: ±4% margin of error. Goldberg, Jeanne (2017). "The Politicization of Scientific Issues: Looking through Galileo's Lens or through the Imaginary Looking Glass". Skeptical Inquirer. 41 (5): 34–39. Archived from the original on August 16, 2018. Retrieved August 16, 2018. Bolsen, Toby; Druckman, James N. (2015). "Counteracting the Politicization of Science". Journal of Communication (65): 746. Freudenberg, William F.; Gramling, Robert; Davidson, Debra J. (2008). "Scientific Certainty Argumentation Methods (SCAMs): Science and the Politics of Doubt" (PDF). Sociological Inquiry. 78: 2–38. doi:10.1111/j.1475-682X.2008.00219.x. Archived (PDF) from the original on November 26, 2020. Retrieved April 12, 2020. van der Linden, Sander; Leiserowitz, Anthony; Rosenthal, Seth; Maibach, Edward (2017). "Inoculating the Public against Misinformation about Climate Change" (PDF). Global Challenges. 1 (2): 1. Bibcode:2017GloCh...100008V. doi:10.1002/gch2.201600008. PMC 6607159. PMID 31565263. Archived (PDF) from the original on April 4, 2020. Retrieved August 25, 2019. External links Look up science in Wiktionary, the free dictionary. vte Philosophy of science Concepts AnalysisAnalytic–synthetic distinctionA priori and a posterioriCausalityCommensurabilityConsilienceConstructCreative synthesisDemarcation problemEmpirical evidenceExplanatory powerFactFalsifiabilityFeminist methodFunctional contextualismIgnoramus et ignorabimusInductive reasoningIntertheoretic reductionInquiryNatureObjectivityObservationParadigmProblem of inductionScientific evidence Evidence-based practiceScientific lawScientific methodScientific pluralismScientific RevolutionScientific theoryTestabilityTheory choiceTheory-ladennessUnderdeterminationUnity of sciencemore... Theories CoherentismConfirmation holismConstructive empiricismConstructive realismConstructivist epistemologyContextualismConventionalismDeductive-nomological modelEpistemological anarchismEvolutionismFallibilismFoundationalismHypothetico-deductive modelInductionismInstrumentalismModel-dependent realismNaturalismPhysicalismPositivism / Reductionism / DeterminismPragmatismRationalism / EmpiricismReceived view / Semantic view of theoriesScientific essentialismScientific formalismScientific realism / Anti-realismScientific skepticismScientismStructuralismUniformitarianismVitalism Philosophy of... BiologyChemistryPhysics Space and timeSocial science ArchaeologyEconomics‎GeographyHistoryLinguisticsPsychology Related topics Criticism of scienceDescriptive scienceEpistemologyFaith and rationalityHard and soft scienceHistory and philosophy of scienceNormative scienceProtosciencePseudoscienceRelationship between religion and scienceRhetoric of scienceScience studiesSociology of scientific ignoranceSociology of scientific knowledge Philosophers of science Precursors Roger BaconFrancis BaconGalileo GalileiIsaac Newton Auguste ComteHenri PoincaréPierre DuhemRudolf SteinerKarl PearsonCharles Sanders PeirceWilhelm WindelbandAlfred North WhiteheadBertrand RussellOtto NeurathC. D. BroadMichael PolanyiHans ReichenbachRudolf CarnapKarl PopperCarl Gustav HempelW. V. O. QuineThomas KuhnImre LakatosPaul FeyerabendIan HackingBas van FraassenLarry Laudan Category Philosophy portalicon Science portal vte Glossaries of science and engineering Aerospace engineeringAgricultureArchaeologyArchitectureArtificial intelligenceAstronomyBiologyBotanyCalculusCell biologyChemistryCivil engineeringClinical researchComputer hardwareComputer scienceDevelopmental and reproductive biologyEcologyEconomicsElectrical and electronics engineeringEngineering A–LM–ZEntomologyEnvironmental scienceGenetics and evolutionary biologyCellular and molecular biology 0–LM–ZGeography A–MN–ZArabic toponymsHebrew toponymsWestern and South AsiaGeologyIchthyologyMachine visionMathematicsMechanical engineeringMedicineMeteorologyMycologyNanotechnologyOrnithologyPhysicsProbability and statisticsPsychiatryQuantum computingRoboticsScientific namingStructural engineeringVirology vte Science and technology studies Economics Economics of scienceEconomics of scientific knowledge History History and philosophy of scienceHistory of science and technologyHistory of technology Philosophy AnthropoceneAntipositivismEmpiricismFuzzy logicNeo-LuddismPhilosophy of sciencePhilosophy of social sciencePhilosophy of technologyPositivismPostpositivismReligion and scienceScientismSocial constructivismSocial epistemologyTranshumanism Sociology Actor–network theorySocial construction of technologyshaping of technologySociology of knowledge scientificSociology of scientific ignoranceSociology of the history of scienceSociotechnologyStrong programme Science studies AntiscienceBibliometricsBoundary-workConsilienceCriticism of scienceDemarcation problemDouble hermeneuticLogologyMapping controversiesMetascienceParadigm shift black swan eventsPseudosciencePsychology of scienceScience citizencommunicationeducationnormalNeo-colonialpost-normalrhetoricwarsScientific communityconsensuscontroversydissententerpriseliteracymethodmisconductpriorityskepticismScientocracyScientometricsTeam scienceTraditional knowledge ecologicalUnity of scienceWomen in science STEM Technology studies Co-productionCyborg anthropologyDesign studiesDematerializationDigital anthropologyDigital media use and mental healthEarly adopterEngineering studiesFinancial technologyHype cycleInnovation diffusiondisruptivelinear modelsystemuserLeapfroggingNormalization process theoryMedia studiesReverse salientSkunkworks projectSociotechnical systemTechnical changeTechnocracyTechnoscience feministTechnological changeconvergencedeterminismrevolutiontransitionsTechnology and societycriticism ofdynamicstheories oftransferWomen in engineering Policy Academic freedomDigital divideEvidence-based policyFactor 10Funding of scienceHorizon scanningPoliticization of scienceRegulation of scienceResearch ethicsRight to scienceScience policy history ofscience ofTechnology assessmentTechnology policyTransition management Portals ScienceHistory of scienceTechnology Category AssociationsJournalsScholars Portal: icon Science Science at Wikipedia's sister projects: Media from Commons News from Wikinews Quotations from Wikiquote Textbooks from Wikibooks Resources from Wikiversity Travel guides from Wikivoyage Categories: ScienceObservationMain topic articles The 20 Most Famous Geniuses in History From the discovery of radium to advances in cancer treatments, see who changed the world as we know it. Andrew Moran Andrew Moran Business and Finance Expert Reviewed by Hayley Ramsey Illustration of Albert Einstein The mind is a glorious vessel that needs to be cultivated and allowed to bloom. Life has become so advanced because millions of the smartest folks throughout history tapped their little gray cells, as the great Hercule Poirot would say, and made our existence better. Indeed, we wouldn’t be where we are today if it were not for some of the top contributions made by these intellectual figures, whether it is in philosophy or science. And, yes, the list goes beyond Elon Musk and Steve Jobs, who typically rank near the top. So, who are these intelligent juggernauts? And how have they helped develop our intelligence? Here is a breakdown of 20 famous geniuses of our time. 1. Baruch Spinoza Would the Age of Reason have ever come about if not for Baruch Spinoza and his colleagues? The Dutch philosopher was instrumental in the foundation of the Age of Enlightenment in the 18th century, contributing to the historical developments in economics, politics and science. Spinoza was integral in the reformation of the Church, challenging the theologians and their status quo. Despite accusations of heresy, Spinoza acknowledged his belief in God, but his view was that this apotheosis was “the sum of the natural and physical laws of the universe and certainly not an individual entity or creator.” 2. Galileo Galilei If it were not for Galileo Galilei, would we be so advanced in the field of physics and astronomy? He is a chief example of someone who changed our world and perspective. The Italian scientist gave us more than just the telescope and the discovery of planetary objects. Galileo was responsible for informing us that we reside in a heliocentric solar system (the Earth revolves around the sun). This was unheard of in the 16th century and prompted an inquisition into what was considered a fanatical and contrarian view to the accepted view of a geocentric solar system (the sun revolves around the Earth). Change the World: Top Jobs to Consider ALSO READ Change the World: Top Jobs to Consider 3. Marie Curie What was remarkable about Marie Curie was that she accomplished a lot of firsts. For instance, she was the first woman to win the Nobel Prize and the first person ever to win two Nobel prizes in two different science fields. But Curie is best known for discovering the elements polonium and radium and helping World War I soldiers with mobile radiography units. She is also revered for her contributions to the advancement of cancer treatment. More specifically, her pioneering research led to the development of radiotherapy for cancer patients. Take our career test Join 500k career hunters and get matched to jobs you'll love and succeed in. TRY IT FREE 4. Leonardo Da Vinci Everyone knows who Leonardo da Vinci is. The quintessential Italian renaissance man needs no introduction. Many of the technological productions that have advanced humankind were conceptualized by none other than Da Vinci. The engineer and artist would put many of our contemporary geniuses to shame! Here are just some examples of inventions that emanated from a man who was ahead of his time: Helicopter Diving suit Parachute Airplane Self-propelled cart Robotics Of course, Da Vinci also delved into the world of art, painting perhaps the most well-known portrait in the history of the world: Mona Lisa. He could do it all. Imagine if he was around today. Maybe finally we could colonize Mars, travel back and forth through time and ensure our shoes smell good. 5. Albert Einstein Like Leonardo Da Vinci, Albert Einstein is one of the most famous geniuses in the history of our world. Unfortunately, his contributions to humanity are not nearly as universal as his name, except his theory of relativity. But Einstein provided us with a treasure trove of information that remains pertinent in the world of mathematics and physics today. In fact, common-day things such as paper towels, stock market forecasts, solar power and laser pointers, exist because of Einstein. 6. Katherine Johnson Katherine Johnson was a mathematician who was instrumental to the first-ever US crewed spaceflights. Thanks to her calculations of orbital mechanics as an employee at NASA, the space agency was able to send people to space successfully. Her work occurred at a time of extreme racial inequality, and her inspiring story and commendable achievements led her to receive a Presidential Medal of Freedom. NASA calls Johnson’s role “historical” as she was one of the first Black American women to take on a scientist position for the space agency. 7. Aristotle Despite being a student of Plato’s, Aristotle was likely far more influential as a Greek philosopher than his teacher. Aristotle had a major influence on the creation and development of Western philosophy, whose ideas played a monumental role in everything, from physics to logic and reason to the theater. He also advanced the area of morality and virtue, presenting the idea that activity of the soul is directly related to moral virtues. Put simply, he argued that practical wisdom was the basis of happiness, which is perhaps one of the all-time important developments in reason. The Weirdest Jobs in the World ALSO READ The Weirdest Jobs in the World 8. Grace Hopper Before there was Steve Jobs or Bill Gates, there was Grace Hopper. Hopper was one of the first computer programmers, pioneering the technology of the first linkers. The United States Navy rear admiral created one of the first working code compilers that spawned COBOL-one, a critical computer programming language of the 20th century. The computer scientist had the guided-missile destroyer USS Hopper named after her, as well as the Cray XE6 “Hopper” supercomputer at the National Energy Research Scientific Computing Center. 9. Jane C Wright Jane Cooke Wright pioneered cancer research and surgery and contributed extensively to chemotherapy. Decades ago, when Wright was just starting out, being diagnosed with cancer was a death sentence. However, because she decided to use human tissue culture instead of laboratory mice to test the effects of possible drugs on cancer cells, Wright forever changed medical cancer treatment. In 1967, when Wright became associate dean and professor of surgery at New York Medical College, she was the highest ranked Black American woman at a nationally recognized medical school in the US. 10. Nikola Tesla If you ask today’s Silicon Valley geniuses who their biggest influences are, the name Nikola Tesla will typically make any relevant list. There is even a revolutionary automobile company named after the man. The Serbian inventor established A/C electricity, wireless transmission of energy, remote control and wireless telegraphy. Tesla was also considered a futurist for envisioning what tomorrow’s world would be like over the next century. Suffice it to say, early writings suggest he had anticipated the prevalence of smartphone and drone technologies. 11. Bill Gates Bill Gates was the driving force for everything we love about the modern-day computer, whether it is a desktop or mobile. With Microsoft at the forefront of technology, Gates gave us the Dos operating system, which is an ancient code today. He then created the Windows operating system, and a brave new world was born. This piece of software changed every aspect of the world, creating a tsunami of new opportunities. Gates’ methods of industry dominance are still questioned and discussed today. Still, his level of intellect that helped satisfy the needs of a billion people allowed him to become one of the richest men on the planet. 12. John Stuart Mill Do you enjoy your freedom? Do you like not being chained to the state? If so, be sure to give John Stuart Mill a read. The English political economist was an imperative source for the development of classical liberalism, which dictated individual freedom. In addition to his written works, he also crafted Mill’s method of agreement, which is utilized by attorneys to this date. Considering that he spent his childhood studying the works of Greek and Latin philosophers and mathematicians, it’s no surprise that Mill became one of the most prominent figures in the shaping of 19th-century political discourse in Britain. 13. Dr Martin Luther King Jr Dr Martin Luther King Jr was a towering figure and will always be remembered as one of the most integral people of the 20th century. The Christian minister and activist remains revered today, and with modern political developments, Dr Luther King Jr’s sage words and vast intellect are more important than ever before. What earned Dr Luther King Jr his Nobel Peace Prize in 1964 was his advocating for non-violent resistance to push for equal rights for Black Americans. The World's Most Influential People ALSO READ The World's Most Influential People 14. Hedy Lamarr Anyone who has ever watched classic films will know Hedy Lamarr as one of the loveliest actresses to grace the silver screen. She starred in some excellent pictures, including Algiers and Ecstasy. But Lamarr was more than just a thespian talent. She was also an inventor and an integral force who contributed to the fundamentals of today’s Wi-Fi, Bluetooth, and GPS communication systems in the 1940s. Lamarr achieved all this despite having been primarily self-taught. Her curiosity around machines and technology apparently became evident at a young age, when, at five years old, she dismantled her music box to make sense of its mechanism. 15. Sir Isaac Newton What Sir Isaac Newton discovered in the 17th century forever changed our world. Most notably, the English scientist’s development of the three laws of motion, his discovery of calculus, and his inventor status in the field of mechanical calculators were imperative contributions to society — then and today. Newton was a key figure in the Scientific Revolution of the 1600s. Indeed, his career was more than just a falling apple! 16. Ludwig von Mises One of the greatest minds in the history of economics was Ludwig von Mises. Everything that he had ever written or uttered remains vital to the Austrian School of Economics, a free-market capitalist philosophy. Despite English not being his first language, he penned numerous books and articles on everything, from economics to political philosophy. His magnum opus that is considered required reading for any economics student is Human Action. Even if you are not interested in this field of study, if you ever wished to delve into the subject, then this is the only book you will never need. 17. Milton Friedman Like Ludwig von Mises, Milton Friedman was also one of the greatest minds in economics. Although he also penned many books, papers and articles, Friedman’s ever-important contribution to the issues of economics and politics was Free to Choose — both his book and the PBS documentary series. He helped form the opinions of a lot of young people, who eventually became leaders in the conservative movement. Even if you disagreed with his ideas, Friedman was still the most charming, approachable and down-to-earth economist you would ever meet. What’s more, the only topic he ever changed his mind about was monetary policy, going from championing the Federal Reserve to advocating its abolition in his later years. 18. Alan Turing The legendary career of Alan Mathison Turing helped become well-known thanks to the 2014 movie The Imitation Game. The film helped the world learn about this British cryptanalyst, computer scientist, logician, mathematician, and theoretical biologist. His life’s work, which was cut short at just the age of 41, led to the modern computer and artificial intelligence. Winston Churchill famously said that Turing’s work and contributions shortened the Second World War by two years. 19. Stephen Hawking When people talk about the world’s top geniuses, Stephen Hawking’s name is usually among the first to come up. The English cosmologist and theoretical physicist is most known for his 1974 discovery that black holes can emit radiation. The phenomenon has since been dubbed “Hawking radiation” and is considered one of the most crucial discoveries about black holes to date. A lesser-known fact about Hawking is that, when he was diagnosed with ALS at the age of 21, his doctors predicted that he would live a few more years at most. However, the cosmologist lived for another five decades. Top Success Stories to Inspire You ALSO READ Top Success Stories to Inspire You 20. Lise Meitner Though you may not have heard of Lise Meitner, you’ve probably heard of “nuclear fission”, a phenomenon she co-discovered in 1938. Nuclear fission, which Meitner was the first to understand, describe and name, refers to the division of the nucleus into smaller nuclei: the principle behind atomic weapons. Outside of this discovery, for which only her collaborator Otto Hahn received a Nobel Prize, the Austrian-Swedish physicist was the second woman to ever earn a doctorate in physics from the University of Vienna. Meitner was also the first woman to become a physics professor in Germany. Final thoughts What has been your latest achievement? While it is always a good idea to start small, it could be time to think bigger and realize your potential. Who knows? A little bit of effort could help you come up with ground-breaking discoveries and you could become the next genius added to the list. Which of these famous geniuses did you find the most inspiring? Share your thoughts in the comments section below! What's your ideal career? Join 500k career hunters and get matched to jobs you'll love and succeed in. TRY THE TEST Originally published July 11, 2020. Updated by Electra Michaelidou. Intelligence Success Stories 7 Comments Andrew Moran Andrew Moran - Business and Finance Expert An experienced business journalist, Andrew specialises in economics and has written for several successful publications, including Liberty Nation, Digital Journal